Figure 7a. Atmospheric transmission, 1 to 5.5 μ ; visual transmission (0.61 μ) 60% per sea mile; water content 17 mm. Figure 7b. Atmospheric transmission, 7.5 to 14 μ; visual transmission (0.61 μ) 60% per sea mile; water content 17 mm.
Figure 4. The 2.7- and 3.2 -μ bands of H2O, showing the absorption parameters variable with the multiple-traversal cell.
Figure 6. Comparison of observed and calculated transmittances in the 18-36 cm-1 region for representative samples of H2O and H2O+N2. The observed values are represented by data points. Calculated values of Tonic (v) with the instrumental response function taken into account are represented by the solid lines, and calculated true transmittance T'(v) by the short-dashed lines, D and F. The long-dashed curves, A, E, and H represent TnIc (v) with the following modifications to the parameters: Panel I, Curve A, the contribution by the continuum was not included. Panel II, Curve E, the absorption coefficient was increased by 10% at all wavenumbers. Panel IV, Curve H, the self-broadening factor Be for the continuum was assumed to be 5 instead of 13. The sample parameters are
The shapes of the extreme wings of self-broadened CO2(lines have been investigated)in three spectral regions near 7000, 3800, and 2400 cm-1. Absorption measurements have been made on the high-wavenumber sides of band heads where much of the absorption by samples at a few atm is due to the extreme wings of strong lines whose centers occur below the band heads. New information has been obtained about the shapes of self-broadened CO2 lines as well as CO2 lines broadened by N2, O2, Ar, He, and H2. Beyond a few cm-1 from the line centers, all of the lines absorb less than Lorentz-shaped lines having the same half-widths. The deviation from the Lorentz shape decreases with increasing wavenumber, from one of the three spectral regions to the next. The absorption by the wings of H2- and He-broadened lines is particularly low, and the absorption decreases with increasing temperature at a rate faster than predicted by existing theories.
Influence of assumed line width on calculated curves of χ (v- vo) and χ (ν) in the 7000 cm-1 region. The curve in the lowerpanel represents the experimental results for ks0. The + 's represent the values calculated on the basis of lines whose χ is given by the solid curve in the upper panel and whose half-widths are given by Fig. 4. The squaries represent values calculated on the basis of the same χ but with α0 = 0.092 cm-1 for all lines. Values of ks0 based on lines with α0= 0.092 cm-1 and x modified according to the dashed curve agree with the experimental curve to within ±42%.
Figure 16. Curves of χ and kN20 showing influence of assumed line shape on the calculated absorption coefficient in the 3800 cm-1 region. Curve A in the lower panel represents the experimental results for KN20. The circles represent the values calculated on the basis of a line shape whose χ is given by curve A in the upper panel. Variations in the line shape given by curves B, C, D, and E in the upper panel were assumed, and the corresponding calculated curves of kN20 are shown in the lower panel. The six vertical lines in the lower panel indicate the wavenumbers at which experimental measurements were made.
The collision‐induced absorption of the (0–0) and (1–0) bands of the 1Δg ← 3Σg− transition of oxygen has been studied in the gas phase from 87°–300°K. Low pressures were used and only binary collisions are important in inducing transitions. At low temperatures the absorption is enhanced because the attractive intermolecular potential causes clustering in the gas. However, no direct evidence for bound‐state (O2)2 molecules was found. The frequency maxima of the absorption of collision pairs were only slightly shifted from those calculated from monomer oxygen. This suggests that the shapes of the internuclear potentials and separation of the 1Δg and 3Σg electronic states are little affected by these binary collisions. A near‐Boltzmann relation of the broad band shapes, moreover, suggests that the intermolecular potentials for 1Δg⋅⋅⋅3Σg−and3Σg−⋅⋅⋅3Σg+ interactions are nearly the same. The mechanisms of the induced absorption collisions are discussed.
Figure 7. Boltmann-weighted Lorentzian curves fitted to the 2 components of the (1-0) 1Δg ← 3Σg-- band at 87° and 300°K. The dashed curves are calculated for the individual components and the solid curve is the sum of these. The curve with dots gives the experimental data.
The collision‐induced absorption spectra have been measured at room temperature and at 87°K for bands in the 1Δg + 1Δg ← 3Σg− + 3Σg− and 1Δg + 1Σg+ ← 3Σg− + 3Σg− simultaneous electronic systems for oxygen. The binary absorption coefficients were found to increase with decreasing temperature for 1Δg + 1Δg ← 3Σg− + 3Σg−. The band shapes for this system suggest that the Hamiltonian which is responsible for intensity borrowing depends on the angular orientation of the O2 molecules in the collision pair since ΔK = 0,± 2,± 4 selection rules are needed to account for the Δν1 / 2 ∼ 200cm−1 bandwidth. The relative intensity of the (1–0) and (0–0) bands indicates that the interaction Hamiltonian is also strongly modulated by the vibrational coordinates of O2. The frequency shift of this simultaneous transition indicates that the intermolecular distance parameter for 1Δg⋅⋅⋅1Δg is 3% larger than for 3Σg−⋅⋅⋅3Σg−. The unusual band shape for the 1Δg + 1Σg+ ← 3Σg− + 3Σg− band is interpreted in terms of an exiton interaction for the 1Δg⋅⋅⋅1Σg>+ combination. Although bound state (O2)2 molecules undoubtedly exist at low temperatures these data provide no unambiguous spectroscopic evidence of their presence.
Figure 1. Spectra of the (0-0) band of the 1Δg + 1Δg ← 3Σg-+3Σg- simultaneous transition. The pathlength is 168 m. The slitwidth is 1 cm-1. The oxygen desnity is 2.1 amagats at both temperatures. The tick mark indicates the band origin calculated from free-molecule spectroscopic constants.
Figure 2. Spectra of the (1-0) band of the 1Δg + 1Δg ← 3Σg-+3Σg-- a simultaneous path length tranSItion. of 122 m The and a upper slitwidth spectrum of about was 1.2 recorded cm-1. with The lower spectrum was recorded with a path length of 168 m and a slitwldth of about 2.5 cm-1. The oxygen density was 2.1 amagats at both temperatures. The tick mark indicates the band origin calculated from the spectroscopic constants of the free molecule.
Figure 1. Влияние температуры на спектральное пропускание паров Н2О (Р0=2.5 атм, w=4*10-3 ос см): а – полоса 1.37 мкм, б – полоса 1.87 мкм; 1 – Т=-30оС, 2 – Т=30оС, 3 – Т=120оС
The two components of the (ν1, 2ν2) Fermi doublet of gaseous CO2 in an absorption path length of 56 m at 192 °K show a complex structure. When the allowed C16O18O absorption and the pressure-induced CO2 absorption are removed by computational procedures, the residual spectrum consists of two similar symmetric patterns of five maxima. These are interpreted in terms of the rotation and vibration of (CO2)2 dimers held by quadrupole–quadrupole interaction in the locked T position at an intermolecular distance of 4.1 Å.
Figure 1. The (v1, 2v2) infrared band of CO2. (a) Ob- served spectrum (I = 56 m, ρ = 0.92 amagat, T = 1920K). (b) Profile with the (v1, 2v2) band of C16O18O removed. (c) Calculated pressure-induced profile for free-free collisions. (d) (= (b) - (c)) Spectrum ascribed to the (CO2)2 dimer. (e) The v1d band with assignments of the maxima.
By means of a photoelectrical technique used in conjunction with a 3 m vacuum monochromator, the absorption cross section of the O2 continuum in the region 2350–1814 Å and the absorption cross section of CO2 in the region 2160–1718 Å have been measured. The cross section of the O2 continuum is 3.8 × 10−24 cm2 at 2350 Å; it slowly increases towards shorter wavelengths and reaches 10.7 × 10−24 cm2 at about 1980 Å, then increases very rapidly and reaches 7.1 × 10−22 cm2 at about 1814 Å. In the case of CO2, numerous discrete bands were found overlapping a weak continuum in the wavelength region below 1980 Å. The absorption cross section of the CO2 continuum is about 2 × 10−24 cm2 at 2100 Å; it gradually increases toward the shorter wavelength side, and reaches about 4 × 10−24 cm2 at 2000 Å. The continuum rises rapidly at 2000 Å and its value is 1.19 × 10−20 cm2 at 1718 Å. Origins of the continua of both O2 and CO2 are briefly discussed.
Table 2. Absorption coefficients and cross sections of the CO2 continuum in the region 2160-1718 A.
Figure 5. Absorption coefficients of CO2 in the region 2160-1880 A. The right and left curves correspond to the right and left scales, respectively.
Figure 6. Absorption coefficients of CO2 in the region 1885-1800 A.
Figure 7. Absorption coefficients of CO2 in the region 1805-1720 A.
The infrared fundamental band and the five strongest near-infrared and visible electronic bands of gaseous oxygen were studied from 90 to 115 K with path lengths up to 140 m in two low-temperature multiple-traversal absorption cells. The profile of the fundamental band is in good agreement with the theory of quadrupole-induced absorption except for a low-intensity residual in the Q-branch region. Although the electronic bands are less amenable to complete analysis, the general validity of a Boltzmann relation in their intensity distributions confirms their collision-induced nature. The temperature variation of the integrated band intensities is indicative of quadrupole induction for the fundamental and of overlap induction for the electronic bands; a somewhat too sharp rise at low temperatures may be due to the neglect of the quadrupole–quadrupole coupling in evaluating the pair distribution function.
Figure. 3. The 1.06 and 1.26 μ bands of O2 at 90 and 112 K. The densities were 2.91 and 4.98 amagat, respectively, and the absorption path was 137 m.
Figure 4. The 5770 and 6290 A bands of O2 at 90, 113, and 295 K. The densities were 2.66, 5.61, and 4.42 amagat, respectively, and the absorption path was 137 m at the low temperatures and 165 m at room temperature.
Figure 1. Кривые прозрачности паров воды при температуре 350оС. Давление, Толщина, Плотность, бар мм г/см3 а 158 0.20 0.1 б 76 0.62 0.032 в 27.5 2.0 0.01 г 8.7 6.2 0.0032 д 2.8 20.0 0.001
Коэффициент поглощения в канте полос 1.4 (а), 2.7 (б) и 4.3 мкм (в): кривые – расчет авторов, точки – эксперимент [5]. The absorption coefficient in the edge of the bands (a) 1.4, (b) 2.7 and (c) 4.3 mm: curves - authors' calculations, points - experiment [5]. [5] Burch D.E., Gryvnak D.A., Patty R.R., Bartky Ch.E. Absorption of infrared radiant energy by CO2 and H2O. IV. Shapes of collision -broadened CO2 lines. J.Opt.Soc Amer. 1969, 59, No.3, 267-280.
Figure 1. Спектры оптической плотности паров воды при температуре 360оС и различных давлениях: а – 65; б – 114; в – 143; г – 177 ата; 1 – наблюдаемый суммарный спектр, 2 – спектр мономерной части молекул, 3 – спектр ассоциированных молекул.
Collision induced microwave absorption is reported in pure CO2, and CO2-Ar, CO2-CH4 mixtures in the 70 G H z (2.3 cm-1) region at a temperature of 22°C, using a sensitive cavity technique previously described. The results in pure CO2 in the very low density region from 5 to 30 amagat accurately establish the dependence of the loss on the square and cube of the density, and the relaxation times are calculated. The experimental results agree well with previously reported lower frequency data at 0.3-0.8 cm-1 which establishes the linear dependence on frequency of the absorption up to 2.3 cm-1. There is also good agreement with an extrapolation of higher frequency infrared results of Ho et al. The relaxation times associated with the two and three body collisions are shown to be nearly equal at room temperatures with T~ = 0.84 x 10-12 s and T~ = 1.0 x 10-12 S. Higher order dependence on the density is observed for the CO2-Ar and CO2-CH4 mixtures. The results are compared with earlier low frequency measurements at 0.8 cm-1 and with the theory of Maryott and Kryder, taking account of correction terms in the dielectric virial coefficient according to Bose and Cole.
Figure 6. A plot of the absorption coefficient a(v) divided by v2 vs. v, showing a solid line through the experimental results for a2(v)/v2 of Ho et al. (1971), Frenkel and Woods (1966). and the present results, and a dotted line through the results of Birnbaum et al. (1971) and the present result for a3(v)/v2.
Спектры поглощения паров воды различной плотности при температуре 350оС. а – 0.0032, б – 0.01, в – 0.032, г – 0.1 г/см3.
Функции пропускания СО2-СО2 и СО2-N2, Т=293 К, L=4.97 м. Сплошные кривые построены по данным эксперимента, пунктирные – по расчетным данным: а) СО2-СО2. 1) р=0.475 атм, 2) р=0.00989 атм; б) СО2-N2, р=0.5 атм
Figure 1. Сопоставление экспериментальных и расчетных функций пропускания. Сплошная кривая – экспериментальные данные, пунктир – расчетные, штрихпунктир – расхождение между ними. 1. u=200 атм см, Рэф=0.889 атм 2. u=500 атм см, Рэф=0.0158 атм [10] Кондратьев К.Я., Тимофеев Ю.М. Численное моделирование функций пропускания для узких спектральных интервалов 15 мкм полосы СО2 // Изв АН СССР, ФАО 1969. Т.5. №4. С.394. Comparison of experimental and calculated transmission functions. Solid curve - experimental data, dotted line - calculated, dash-dotted line - discrepancy between them. 1.u = 200 atm cm, Ref = 0.889 atm 2.u = 500 atm cm, Ref = 0.0158 atm [10] Kondrat'ev K.Ya., Timofeev Yu.M. Numerical modeling of transmission functions for narrow spectral intervals of 15 µm of the CO2 band // Izv AN SSSR, FAO 1969. V.5. No. 4. P.394.
Рисунок 1(не полный). Влияние температуры на спектры поглощения в полосах 2.0 и 2.7 мкм углекислого газа. Figure 1(reduced). Influence of temperature on absorption spectra in the 2.0 and 2.7 μm bands of carbon dioxide.
Figure 1. Спектры пропускания водяного пара в области 1.35 (а) и 1.87 (б) мкм: а : 1 – w=0.084 г/см2, РН2О=4.2 атм, РН2О-N2=120 атм, 2 - w=0.0083 г/см2, РН2О=41.44 атм, 3 - w=0.0098 г/см2, РН2О=4.954 атм, пунктир – спектр димера б : 2 - w=0.0086 г/см2, РН2О=4.355 атм, 1 - w=0.0083 г/см2, РН2О=41.44 атм, 3 - w=0.0084 г/см2, РН2О=4.2 атм, РН2О-N2=120 атм.
In this paper we report the results of a series of pressure-broadened water vapor absorption measurements at 27 CO2 laser frequencies between 935 cm-1 and 1082 cm-1. Both multiple traversal cell and optoacoustic (spectrophone) techniques were utilized together with an electronically stabilized cw CO2 laser. Comparison of the results obtained by these two methods shows remarkable agreement, indicating a precision which has not been previously achieved in pressure-broadened studies of water vapor. The data of 10.59 Am substantiate the existence of the large (>200) self-broadening coefficients determined in an earlier study by McCoy. In this work we have treated the case of water vapor in N2 at a total pressure of 1 atm. We have also studied water vapor in air and will report those results separately.
Table 1. List of Spectrophone and White Cell Curve Fit Coefficients for H2O in N2. Measurements at a Total Pressure of 760 Torr. The data have been fit to an equation of the form k(v) = C°s(ν) WH2O(PH2O +γ (P - PH2O)). Defined function = γ(ω). Dependence of the gamma coefficient on the wavenumber in above formula. Note: In original article the coefficient γ is ansent in the formula.
Figure 1. Curves a-a’ and b-b’ show observed anomalous absorption values for mean temperatures of 281° and 290°K respectively, reduced to a water vapour density of 4.8 g m-3. The estimated error is ±0.02 dB km-1. Curves A-A’ and B-B’ are the corresponding temperature dependence values expressed as the exponent B in equation (1), with a maximum estimated error ±0.04 eV per molecule (at 12.6 μm wavelength). Curve l-l’ shows laboratory values from Ref.3 scaled to 290°K and density 4.8 g m-3, and L-L’ is the corresponding laboratory temperature dependence. Curve m-m’ shows a monomer model spectrum from the data in Table 1, for 290°K and density 4.8 g m-3. The values given by P and Q are taken from Ref.1, and apply to the temperature range 258° to 299°K.
[1]. Coffey, M.T., Quat. Jour. Res. Met. Soc, 103, 685-692 (177) [3] Burch D.E., Proc. Am. Met. Soc, Conference on Atmospheric Radiation, Fort Collins, Colorado, 7-9 August (1972).
Figure 1. Случай самоуширения. РСО2=1 атм, Т=300°К, эксперимент [2], приближение сильной линии, сумма по линиям. Self-broadening case. РСО2=1 atm, Т=300°K, experiment [2], approximation of a strong line, the sum of the lines. а) 4.3 μm, m=20, C20=1.3 10-10 cm-1; b) 2.7 μm, m=20, C20=1.8 10-10 cm-1; c) 1.4 μm, m=20, C20=2.5 10-10 cm-1. [2] Burch D.E., Gryvnak D.A., Patty R.R., Bartky Ch.E. Absorption of infrared radiant energy by CO2 and H2O. IV. Shapes of collision -broadened CO2 lines. J.Opt.Soc Amer. 1969, 59, No.3, 267-280
Figure 2. Случай уширения азотом N2. РN2=1 атм, Т=300°К, эксперимент [2], приближение сильной линии, полинейный расчет. N2-broadening case. РN2 = 1 atm, Т=300°K, experiment [2], approximation of a strong line, iine by line calculation. а) 4.3 μm, m=30, C30=1.42 10-17 cm-1; b) 2.7 μm, m=30, C30=1.8 10-17 cm-1; c) 1.4 μm, m=30, C30=5.0 10-18 cm-1; [2] Burch D.E., Gryvnak D.A., Patty R.R., Bartky Ch.E. Absorption of infrared radiant energy by CO2 and H2O. IV. Shapes of collision-broadened CO2 lines. J.Opt.Soc Amer. 1969, 59, No.3, 267-280.
Figure 1. Theoretical water vapor absorption (water vapor density 18 g/m3, temperature 296°K) based on the Gross line-shape formula with laser frequencies in the gaps shown by arrows.
Figure 2. Отклонения рассчитанных K2(νk) от экспериментальных. 1 – χ(ν)=1, 2 – χ(ν) описана кривой (8). СO2+N2, СO2+Ar, СO2+He. Deviations of the calculated K2 (νk) from the experimental ones. 1 - χ (ν) = 1, 2 - χ (ν) is described by curve (8). CO2-He. χ(ν) = x5/2K5/2(x)(1+αx2), x=| ν - νj |/Δ (8), where K5/2(x) is a modified Bessel function. Parameters α and Δ are determined from experimental data.
Зависимость коэффициента поглощения от частоты в случае СО2+СО2 для разных полос. а – 4.3, б – 2.7, в – 1.4 мкм; 1 – эксперимент [4], 2 – расчет в «приближении одной сильной линии», 3 – полинейный расчет, L – расчет по лорентцевскому контуру. Frequency dependence of the absorption coefficient in the case of CO2+CO2 for different bands. a - 4.3, b - 2.7, c - 1.4 μm; 1 - experiment [4], 2 - calculation in the "approximation of one strong line", 3 - linear calculation, L - using Lorentzian contour calculation. [4] Burch D.E., Gryvnak D.A., Patty R.R., Bartky Ch.E. Absorption of infrared radiant energy by CO2 and H2O. IV. Shapes of collision -broadened CO2 lines. J.Opt.Soc Amer. 1969, 59, No.3, 267-280.
Зависимость коэффициента поглощения от частоты в случае СО2+N2 для разных полос. а – 4.3, б – 2.7, в – 1.4 мкм; 1 – эксперимент [4], 2 – расчет в «приближении одной сильной линии», 3 – полинейный расчет, L – расчет по лорентцевскому контуру. Frequency dependence of the absorption coefficient for CO2 + N2 for different bands. a - 4.3, b - 2.7, c - 1.4 μm; 1 - experiment [4], 2 - calculation in the "approximation of one strong line", 3 - linear calculation, L - Lorentzian contour calculation. [4] Burch D.E., Gryvnak D.A., Patty R.R., Bartky Ch.E. Absorption of infrared radiant energy by CO2 and H2O. IV. Shapes of collision -broadened CO2 lines. J.Opt.Soc Amer. 1969, 59, No.3, 267-280.
Зависимость коэффициента поглощения от частоты в случае СО2+СО2 для полосы 1.4 мкм; 1 – эксперимент [4], 2 – расчет в «приближении одной сильной линии», 3 – полинейный расчет, L – расчет по лорентцевскому контуру.
[4] Burch D.E., Gryvnak D.A., Patty R.R., Bartky Ch.E. Absorption of infrared radiant energy by CO2 and H2O. IV. Shapes of collision -broadened CO2 lines. J.Opt.Soc Amer. 1969, 59, No.3, 267-280.
Figure 1. Сравнение экспериментов [6] (а, Н2О-Н2О) и [1] (с учетом [6,7]) (б, Н2О-N2) и расчетов. Аппроксимация экспериментов (1), расчеты по обобщенному контуру (2), по формуле Лоренца (3) и полной формуле Лоренца (4).
[1] Bignell K.J., Saiedy F., Sheppard P.A. On the atmospheric infrared continuum , J. Opt. Soc.America 53, No.4, 466-479 (1963) [6] Burch D.E. Investigation of the absorption of infrared radiation by atmospheric gases. Semi-annual Technical report. Air Force Cambridge Research Lab., Publ. U-4784 (1970) [7] K. J. Bignell, The water-vapour infra-red continuum, Quarterly Journal of the Royal Meteorological Society, Volume 96 Issue 409, Pages 390 - 403 1970 10.1002/qj.49709640904.
The continuum absorption by H2O has several characteristics that are common throughout the windows in the infrared and millimeter-wave regions. Values of the continuum absorption coefficient calculated on the basis of simple, widely used line shapes may differ greatly from observed values in the windows between strong absorption lines. The temperature dependence of this absorption is also not predictable from present day understanding of line shapes or of dieters, which may also contribute. The shapes of self-broadened H2O lines are quite different from those of N2-broadened lines, and the difference increases with increasing distance from the centers of the lines. Data obtained from laboratory samples and from atmospheric paths are presented to compare the various windows in the infrared and millimeter regions.
Figure 13. Comparison of the spectral curves from 0 to 3100 cm-1 of the empirical continuum for self broadening (C), the absorption coefficient of liquid water (L), and the average intensities of the H2O vapor lines (V). Note scale change at 50 cm-1.
Таблица 1. Бинарные коэффициенты поглощения (см-1амага-2) для полосы 3ν3 СО2. Table 1. Binary absorption coefficients (cm-1amaga-2) for the 3ν3 CO2 band.
Рисунок 1. Коэффициент поглощения СО2 в районе 1.4 мкм. CO2 absorption coefficient in the region of 1.4 μm. [1] Burch D.E., Gryvnak D.A., Patty R.R., Bartky Ch.E. Absorption of infrared radiant energy by CO2 and H2O. IV. Shapes of collision -broadened CO2 lines. J.Opt.Soc Amer. 1969, 59, No.3, 267-280. [2] Baranov Yu.I., Bulanin M.O., Tonkov M.V. Study of the wings of the lines of the vibrational-rotational band 3ν3 СО2. Optics and Spectroscopy, 50, no. 3, p. 613-615 (1981).
Figure 1. Predicted spectra for fogs comprising molecular absorption and particle loss computed from standard models; corresponding to visibilities of 50 (dotted line), 10 (dashed line) and 150 (solid line) m. Spectra observed in the atmosphere corresponding to the models. The standard deviation in the measured spectra is 3 dB km-1. In such fogs the liquid drops are all small compared with the wavelengths and show Rayleigh region behavior in which attenuation varies monotonically. Nether this component nor equilibrium dimers can accounr for the observed additional structure between monomer lines.
Absorptions of the ρστ, 0.8 μm and a bands of water vapor were measured using a long pathlength, multiple-traversal absorption cell. The band intensities were found to be 842.3, 53.2, and 50.12 cm-1/g-cm-2 at 334 K for the ρστ, 0.8 μm and a bands, respectively. The measured equivalent widths gradually exceed calculated values for the Lorentz line profile and the self-broadening coefficients of 5 with increasing absorber amounts. Possible causes of this excess absorption are discussed.
Figure 1. Predicted spectra for fogs comprising molecular absorption and particle loss computed from standard models; corresponding to visibilities of 50(…), 10(---) and 150(___) m. Spectra observed in the atmosphere corresponding to the models. The standard deviation in the measured spectra is 3 dB km-1. In such fogs the liquid drops are all small compared with the wavelengths and show Rayleigh region behavior in which attenuation varies monotonically. Nether this component nor equilibrium dimers can accounr for the observed additional structure between monomer lines.
Absorption of 9.6-µm CO2 laser radiation by CO2 at temperatures between 296 and 625 K has been measured at a pressure of 200 Torr. Experimental results for the R10—R26 and P10—P28 transitions have been obtained and compared with computed values of absorption. The relative optical broadening coefficients due to He and N2 have been measured on the R16—R22 and P16—P22 transitions over the same temperature range.
Figure 1. Experimental and calculated values of absorption coefficientk at T = 620°K over the R and P branches of the 9.6-μm CO2 transition.
Using a tunable source of monochromatic radiation (BWO) and a pneumatic detector, the field and laboratory investigations have been performed to obtain spectral distribution of atmospheric water vapor absorption coefficient in transparent windows centered at the wavelengths of 0.88 and 0.73 mm.
Figure 1. Results of our laboratory measurements in comparison with laboratory investigations of (30) and theoretical data (ρ=19 g.m-3, T=25.5oC, P =730 Torr). Our total absorption (A) and excess one (B) are shown by circles. Theoretical values for water vapor monomers (chain-dot curve), excess absorption spectrum (30) (solid curve) and theoretical dimer absorption (7) (dotted curve) are shown as well.
Figure 2. Our results of field measurements as compared with field investigation (30) and the theory (r =8.5 g.m-3, T=8.5°C, P= 727 Torr). Our total absorption (A) and excess one (B) are shown by circles. All the data are shown as in Figure 1. (A) and excess one (B) are shown by circles. Theoretical values for water vapor monomers (chain-dot curve),excess absorption spectrum (30) (solid curve) and theoretical dimer absorption (7) (dotted curve) are shown as well.
[7]. A.A. Viktorova, S.A. Zhevakin, Поглощение микрорадиоволн димерами водяного пара атмосферы, DAN USSR, v.194, 540 (1970). [30]. R.J. Emery, P. Moffat, R.A. Bohlander, H.A. Gebbie, Measurements of anomalous atmospheric absorption in the wavenumber range 4-15 per cm, J. Atm. Terr. Phys. v.7,587 (1975).
Figure 2. Normalized absorption coefficient B0(σ,T) in cm-1amagat-2 for CO2 broadened by N2: (a) wave number dependence of B0(σ,T) for two temperatures 296° and 193°K; (b) temperature dependence at various wave numbers (cm-1).
Figure 3. Normalized absorption coefficient B0(σ,T), in cm-1 amagat-2 for CO2 broadened by O2: (a) wave number dependent of B0(σ,T) for two temperatures 296° and 193°K; (b) temperature dependence at various wave numbers (cm-1).
The collision-induced absorption (CIA) spectrum for nitrogen has been measured in the spectral region below 360 cm−1 at 126, 149, 179, and 212 K. The measurements have been obtained using Fourier transform infrared (FTIR) techniques, a far infrared (FIR) laser system operating at 84.2 and 15.1 cm−1, and microwave cavity techniques. The experimental line shapes have been compared with the theoretical predictions of Joslin, based on Mori theory, and of Joslin and Gray, based on information theory alone. The data have been used to determine the quadrupole moment employing various intermolecular potentials. One Lennard–Jones potential has resulted in a quadrupole moment of 1.51 B, the value that was used in generating the theoretical line shapes. These results, when combined with our forthcoming measurements on nitrogen mixed with methane and argon, may be helpful in determining the role of CIA in calculating the opacity of some planetary atmospheres.
Figure 3. A plot of A(v)/ρ2 versus the frequency (cm-1) measured using an FIR interferometer and indicated by the solid curve. The results are displayed for the four different temperatures used: (a) 212 K, (b) 179 K, (c) 149 K, (d) 126 K. In each case the laser measurements at 84.2 and 15.1 cm-1 are shown. Also shown are the theoretical line shapes generated using Mori theory (dashed line) as discussed in the text. In (d) the results of Buontempo et al. (10) are given for comparison.
A practical atmospheric Millimeter‐Wave Propagation Model (MPM) is formulated that predicts attenuation, delay, and noise properties of moist air for frequencies up to 1000 GHz. Input variables are height distributions (0–30 km) of pressure, temperature, humidity, and suspended droplet concentration along an anticipated radio path. Spectroscopic data consist of more than 450 parameters describing local O2 and H2O absorption lines complemented by continuum spectra for dry air, water vapor, and hydrosols. For a model (MPM*) limited to frequencies below 300 GHz, the number of spectroscopic parameters can be reduced to less than 200. Recent laboratory measurements by us at 138 GHz of absolute attenuation rates for simulated air with water vapor pressures up to saturation allow the formulation of an improved, though empirical water vapor continuum. Model predictions are compared with selected (2.5–430 GHz) data from both laboratory and field experiments. In general, good agreement is obtained.
Figure 8. Water vapor attenuation rates α(v) across atmospheric window range W4 at two temperatures, 5o and -10o C; pluses, measured data [Fedoseev and Koukin, 1984]; lines, MRM.
Figure 1. Absorption coefficients A(V) at 195° (upper) and 296°K (lower). ooooo present results;---- from Birnbaum (1975). A(v) are plotted in arbitrary units. In our data the maximum value is 9.24 10-6 cm-1 amagat-2 at 296°K and 1.22 10-5 cm-1 amagat-2 at 195°K.
Birnbaum, G. Far infrared collision‐induced spectrum in gaseous methane. I. Band shape and temperature dependence J. Chem. Phys., 62:59, 1975, https://doi.org/10.1063/1.430239.
The temperature dependence of the high frequency far wings of the self-broadened CO2 lines has been investigated in the 2400–2600-cm-1 spectral region. The temperature dependence of the corrective shape factor X(σ,T) is demonstrated for the first time.
Figure 5. Wave number dependence of the normalized absorption coefficient A0(σ,T), in cm-1 amagat-2, for two temperatures: 296° and 193°K.
Figure 9. Comparison of experimental and calculated spectrum (A0 in cm-1 amagat-2) (a) T=296°K, (b) T = 218°K: *, observed, ----, Lorentz absorption, -, best fit obtained with the two-parameter lineshape factor of Birnbaum [see Eq. (5)]. The optimized values of the parameters are given in Table IV.
Defined function = log10A0 + 10
Figure 4. Коэффициент поглощения κ(ω) в крыле полосы 4.3 мкм СО2 при высоких температурах. А). Расчет с V(T0=293°K) при Т=300°, 473°, 673°К, соответственно. Б). Расчет с V(T) при тех же температурах. Figure 4. The absorption coefficient κ (ω) in the wing of the 4.3 µm CO2 band at high temperatures. A). Calculation with V (T0 = 293°K) at T = 300°, 473°, 673°K, respectively. B). Calculation with V (T) at the same temperatures.
Figure 5. Коэффициент поглощения СО2 при самоуширении в микроокнах и за кантом полосы 4.3 мкм при различных температурах. Эксперимент [22,3] при Т=300°, 213° К, соответственно. Расчет для Т=213°, 310°, 500°К, соответственно; расчет с F(R)=1. a) Расчет с V(T0); b) Расчет с V(T); c) Расчет с дисперсионным контуром CO2 absorption coefficient for self-broadening in micro-windows and behind the band edge is 4.3 µm at different temperatures. o, + - experiment [22,3] at T = 300°, 213°K, respectively. Calculation for T = 213°, 310°, 500°K, respectively; Calculation with F(R)=1. a) Calculation with V (T0); b) Calculation with V(T); c) Calculation with a dispersion contour. [22] Winters B.H., S. Silverman, W.S. Benedict Line shape in the wing beyond the band head of the 4.3 μ band of CO2 JQSRT V. 4, Issue 4, 1964, Pages 527-537. [3] Буланин М.О., Булычев В.П., Гранский П.В., Коузов А.П., Тонков М.В. Исследование функций пропускания СО2 в области полос 4.3 и 15 мкм. В кн.: Проблемы физики атмосферы. Вып. 13, Л., Изд. ЛГУ, 1976, с.14-24.
Figure.1. Comparison of measured (Ref.5) (dots) and theoretical (curves) roto-translational spectra of H2-CH4 pairs at five temperatures from 140 to 296 K and frequency range from 150 to 850 cm-1.
Existing laboratory measurements of the far-infrared collision-induced spectra of gaseous nitrogen at temperatures from 124 to 300 K are analyzed on the basis of quantum line shapes computed from a suitable, advanced isotropic potential and multipole-induced dipole functions. The input is chosen to represent most closely the measurements at all temperatures and over the full range of frequencies. Simple analytical expressions are specified which represent the spectral profiles closely. It is thus possible to reproduce the collision-induced absorption spectra of nitrogen effortlessly in seconds at temperatures from 50 to 300 K on small computers, even in the far wings which never have been modeled from a quantum formalism before. The work thus gives new and reliable spectral intensities and their temperature dependence for a detailed analysis of the Voyager IRIS spectra of Titan's atmosphere.
Figure 2. Far-infrared absorption spectrum of N2+N2 at (a) 126, 149, and 179 K and (b) 228.3° and 300°K. Open circles, experimental data at (a) 124°K, and (b) 300°K of Buentempo et.al. (1975). Filled circles, squaries and triangles denote new results (Stone et al. 1984; Dagg et al. 1985). Lines, fitted spectrum, with the Ling and Rigby isotropic potential (see Appendix) and the dipole parameters given in Table 3. In (a), the scale is shifted by a factor of 2 at each temperature.
U. Buontempo, S. Cunsolo, G. Jacucci, and J. J. Weis, The far infrared absorption spectrum of N2 in the gas and liquid phases, The Journal of Chemical Physics, 1975, Volume 63, Pages 2570, DOI: 10.1063/1.431648, https://doi.org/10.1063/1.431648 Dagg, I.R., and Gray C.G., 1985, in Phenomenoa induced by intermolecular interaction, ed. G.Birnbaum (New York; Plenum), p.109. Dagg, I.R., Anderson, A., Yan, S., Smith, W. and Read, L.A.A., Collision-induced absorption in nitrogen at low temperatures, Canadian Journal of Physics, 1985, Volume 63, Issue 5, Pages 625-631, DOI: 10.1139/p85-096, https://doi.org/10.1139/p85-096. Stone, N. W. B., Read, L. A. A., Anderson, A., Dagg, I. R., & Smith, W., Temperature dependent collision-induced absorption in nitrogen, Canadian journal of physics, 62(4), 338-347. (1984).
Figure 1. Frequency dependence of the normalized absorption coefficient BN20 (σ,T) (in cm-1 amagat-2) in the centers of the R-branch troughs for N2 broadening: 296°K; 238°K; 193°K.
Figure 2. Frequency dependence of the normalized absorption coefficient BO20 (σ,T) (in cm-1 amagat-2) in the centers of the R-branch troughs for O2 broadening: 296°K; 238°K; 193°K.
Figure 1. Comparison of measured (dots) and theoretical (curves) roto-translational spectra of H2+N2 pairs at five temperatures from 91° to 298°K.
Figure 4. Plots of the absorption coefficient per product density (amix) as a function of the frequency v at the four temperatures (a) 212°, (b) 179°, (c) 149°, and (d) 126°K. The experimental results are shown as the solid curves in the figures, and our theoretically derived line shapes are represented as the dotted curves. In (d), the theoretical contributions due to the various induction mechanisms are labelled Q, 0, and H for quadrupole, octopole, and hexadecapole induction, respectively.
The rototranslational absorption spectrum of gaseous methane has been measured at seven different temperatures from 296 to 140 K. We have analyzed both the spectral moments and the experimental absorption shapes, assuming that only octupolar and hexadecapolar induction mechanisms contribute to the absorption. This assumption allows us to parameterize the temperature dependence of both the intensity and the shape of the absorption band. The results obtained indicate that other contributions to absorption are not negligible.
Experimentally determined absorption band at 163°K and the best-fit curve obtained by using the MLEW model to describe the single line profiles. (a) octupolar contribution; (b) hexadecapolar contribution.
Figure 3. Absorption coefficient AHM(v ) at 296° and 195°K. The vertical bars indicate the typical uncertainties in the experimental points. The broken line is the CH4+H2 absorption coefficient as obtained in previous measurements
Figure 1. Коэффициент поглощения κ(ω) в крыле полосы 4.3 мкм СО2 при высоких температурах. Расчет с V(T0=293°K) при Т=300°, 473°, 673°К, соответственно. Расчет с V(T) при тех же температурах. Absorption coefficient κ (ω) in the wing of the 4.3 µm CO2 band at high temperatures. Calculation with V (T0 = 293°K) at T = 300°, 473°, 673°K, respectively. Calculation with V (T) at the same temperatures.
Existing rototranslational collision-induced absorption (CIA) spectra of methane pairs are analyzed with the help of spectral profiles computed from quantum mechanics. Dipoles induced by octopolar and hexadecapolar fields, hexadecapolar overlap, and the gradient of the octopolar field are considered. The spectral contributions of both bound and free pairs of molecules are accounted for. The analysis which suggests a centrifugal distortion of rotating methane molecules permits one to reproduce from theory the measured CIA spectra at all temperatures (126–300 K) and over the full range of frequencies (> 700 cm−1 at high temperatures) with rms deviations that are smaller than the experimental uncertainties. The values of the octopole and hexadecapole moments of (nonrotating) CH4 molecules needed for that purpose are consistent with state-of-the-art ab initio computations for the first time in such work.
Figure 1. Absorption coefficients α(ω) of CH4-CH4 fitted at different temperatures. Dots: experimental data at temperatures from 140° and 295°K from (20); and at 126°K from (22). J-dependent multipole moments according to Eqs. (9) and (10) are assumed (Table III). For all temperatures above 126°K the displayed spectra are shifted up by one decade; the maximum of the absorption coefficient at each temperature is near 10-5 cm-1 amagat-2 at all temperatures. As an example, the contributions due to the octopole and hexadecapole induction are shown for one temperature.
20. P. Codastefano, P. Dore, and L. Nencini, J . Quant. Spectrosc. Radiat. Transfer 35, 255-263 (1986). 22. I. R. Dagg, A. Anderson, S. Yan, W. Smith, C. G. Joslin, and L. A. A. Read, Canad. J. Phys.1986, in press.
In previous work (Borysow and Frommhold, 1986) rigorous quantum computations of the rototranslational absorption spectra of CH4-CH4 pairs have been communicated which closely reproduce existing laboratory measurements at temperatures from 124 to 300 K and over a range of frequencies from 0 to 750 per cm. Since the computations are complex, this work shows how the spectra can be reproduced from simple, analytical functions that closely model the quantum profiles and, thus, the laboratory measurements. For temperatures from 50 to 300 K, the rototranslational, collision-induced spectra of methane pairs can thus be obtained accurately on small computers in seconds. No measurements exist for comparison with the computational results at the lower temperatures (less than 124 K) but these data are believed to constitute the most dependable theoretical prediction that can presently be made.
Figure 1. CIA spectra of CH4-CH4 at the temperatures of 50°, 110°, and 300°K, computed from a quantum mechanical formalism described elsewhere (Borysow and Frommhold 1987). The dimer bands are convoluted with a triangular slit function of 4.3 cm-1 resolution. The octopole-induced (dashed curve) and hexadecapole-induced (dot-dashed curve) contributions are shown, together with the total (solid curve).
The rototranslational absorption spectrum of gaseous N2 is analyzed, considering quadrupolar and hexadecapolar induction mechanisms. The available experimental data are accounted for by using a line-shape analysis in which empirical profiles describe the single-line translational profiles. We thus derive the simple procedure that allows one to predict the N2 spectrum at any temperature. On the basis of the results obtained for the pure gas, we also propose a procedure to compute the far-infrared spectrum of the N2–Ar gaseous mixture. The good agreement between computed and experimental N2–Ar data indicates that it is possible to predict the far-infrared absorption induced by N2 on the isotropic polarizability of any interacting partner.
Figure 1. Pure N2 absorption spectrum at 297 and 149 K: ... , experimental data from ref. 7; ----. best fit; - - -, hexadecapolar component. Vertical bars indicate typical uncertainties for the experimental points. At 297K, the best-fit parameters are SΦnn = 1690 K-Å6, δ0nn = 19.6 cm-1, S0nn = 49 K-A", and δΦnn = 34.5 cm-1. At 149 K, s:, = 1915 K-Å6, δ0nn = 15 cm-1: SΦnn = 47 K-Å6, and δΦnn = 23.5 cm-1.
Figure 2. Pure N2 absorption spectra at 140 and 93 K: ... , experimental data from ref. 9; ----, computed spectra multiplied by a normalizing factor F (F = 0.93 at 140 K and 1.12 at 93 K); ---, computed 90-K spectrum from ref. 5 (multiplied by a factor of 2 because of a different definition of the absorption coefficient).
A spectrum of the carbon dioxide trimer van der Waals species has been recorded near 3614 cm−1 at sub‐Doppler resolution using an optothermal (bolometer‐detected) molecular‐beam color‐center laser spectrometer. A planar, cyclic structure with C3h symmetry has been determined for the complex with a carbon–carbon separation of 4.0382(3) Å. The observed perpendicular band, corresponding to an in‐plane E′‐symmetry vibration of the trimer, has been attributed to a localized excitation of the 2ν02 +ν3 combination mode of a CO2 subunit by virtue of its small blue shift (∼0.98 cm−1) from that of the isolated monomer.
Figure 1. Observed and calculated spectra of the CO2 trimer in the region of the 2ν02 + ν3 combination band of the monomer indicated by the negative-going R (0) line. Observed spectra for 3% CO2:He mixture at 2, 3, and 5 atm driving pressure. Asterisks indicate transitions attributed to the CO2 dimer. Calculated transitions are simulated with Gaussian profiles of 20 MHz FWHM and an effective rotational temperature of 1.3 K
This is a comprehensive study of water-vapor line and continuum absorption in the 8–14μm atmospheric window by laser-photoacoustic spectroscopy. The characteristics of laser-photoacoustic spectroscopy and detectors are discussed with results on continuum and line absorption at selected CO2-laser wavelengths.
We have assigned four weak absorption lines which occur at the CO2-laser emissions 10P(40), 10R(20), 9P(38) and 9R(36) to pure rotational transitions of H2O, and have determined the dependence of the continuum water-vapor absorption over the temperature range +70°C and −20°C. The measured negative temperature coefficient of the continuum is consistent with both monomer and dimer models, yet not with predictions of larger water clusters.
Experiments with supersaturated water vapor indicate that for S ⩾ 1 collision broadening of distant strong lines as well as water dimer absorption contribute to the continuum. However, the dimer absorption is an order of magnitude too small to cause a significant contribution at ambient atmospheric conditions.
We have investigated the effect of UV-radiation on the 8–14μm absorption of water vapor, buffered either with N2 or synthetic air. The observed changes are explained by UV-photodissociation of H2O molecules and by ozone production. There is no evidence in favor of a cluster model.
Finally, we compared our measured spectra with LOWTRAN 6 and HITRAN models. The LOWTRAN yields a stronger negative temperature dependence than observed while the HITRAN does not predict the observed continuum absorption.
Figure 19. Measured temperature dependence of the pressure quadratic coefficients Cs of the continuum absorption at the (a) 10P(20), (b) 10P(24) and (c) 10P(30) CO2-laser emissions. Dashed lines correspond to least square fits based on equation (25). Solid lines correspond to best fits on the basis of equation (29).
Dimer model: Suck S.H., Kassner J.L., Jr., Jamaguchi J. Water clusters interpretation of IR absorption spectra in the 8-14 μm wavelength region Appl.Opt. 18, No.15, 2609-2617 (1979).
Figure 20. Measured temperature dependence of the pressure quadratic coefficients Cs of the continuum absorption at the (a) 10P(20), (b) 10P(24), (c) 10P(30) and (d) 10P(38) CO2-laser emissions. Solid lines represent best fits on the basis of equation (29).
Figure 24. Temperature dependence of the pressure quadratic coefficients Cs of the continuum absorption for the (a) 10P(20) and (b) 10P(24) CO2-laser emissions. Solid lines represent best fits on the basis of equations (26) and (29).
An F‐center laser–molecular beam spectrometer has been used to obtain a sub‐Doppler resolution infrared spectrum of the carbon dioxide dimer. The vibrational mode investigated in this study corresponds to the ν1+ν3 combination mode of the monomer located at 3716 cm−1. A qualitative assignment of the spectrum shows unambiguously that the equilibrium structure of the dimer is the slipped parallel, rather than the T‐shaped, geometry. The observed spectrum cannot be fit to within experimental error using conventional asymmetric rotor formalism. This may be due to a number of factors such as Fermi resonance between the upper state levels of the band and nearby levels of the dimer, such as seen in the monomer, or it could arise from tunneling effects arising from the two large amplitude motions in the dimer.
Figure 3. A comparison between a calculated and experimental carbon dioxide dimer spectrum. The calculated spectrum was obtained using the rotational constants given in Table II. Although most of the features in the spectrum are properly accounted for there are some extra transitions in the observed spectrum, most likely associated with the parallel band. A very accurate fit to the spectrum was not possible due to what appears to be a staggering of the K band origins due to tunneling.
Figure 2. The collision-induced spectrum of pure N2 + Ar at 126, 149, 179, 212, and 298°K. The measurements are from Dagget al. (1986). Points, FTIR data: Δ. microwave data at 15.1 cm-1; O, laser data at 84.2 cm-1). The solid lines were calculated with the semi-empirical model described in the text. The spectra for different temperatures are vertically offset.
Figure 5. The collision-induced spectrum of N2 + H2 mixture at 91, 141, 165, 195 and 298°K. The measurements are from Dore et al. (1986).
Figure 3. The collision-induced spectrum of pure CH4 at 140°, 163°, 195°, and 295°K. The measurements are from Codastefano et al. (1985, 1986).
Figure 4. The collision-induced spectrum of N2 + CH4 mixture at 126°, 149°, 179°, and 212°K. The measurements are from Dagg et al. (1986).
Existing measurements of the collision-induced rototranslational absorption spectrum of gaseous mixtures of helium and methane are analyzed using quantum profiles computed from recent interaction potentials and various assumptions concerning the interaction-induced dipole components. The computed profiles are expressed in terms of simple functions that reproduce closely both the computed spectra at temperatures from 40 to 350 K and the existing measurements at temperatures from 150 to 353 K. Such a simple analytical representation of the HeCH4 rototranslational spectra facilitates accurate modeling of planetary atmospheres in the far-infrared region of the spectrum.
Figure 1. Density normalized absorption coefficients α(ω)/n1n2 of He+CH4 at three temperatures. Dots: Experimental data at 150°, 293°, and 353°K by Afanas’ev et al. (7). Solid line: Calculated spectra determined by fitting the range and amplitude of only the isotropic (A = 0) overlap contribution to the dipole induction. For all the temperatures above 150°K the displayed spectra are shitkl up by one decade; the maximum of the absorption coefficient at each temperature is near 10-6 cm-1/amagat2.
7. A. D. Afana'sev, M. 0. Bulanin, and M. V. Tonkov, Sov. Technol. Phys. Lett. 6, 1444-1446 (1980). [in Russian]
Figure 2. Коэффициент поглощения крыльями линий СО2: a – самоуширение, б – уширение азотом: I- расчет с теоретическим контуром, II – расчет с уточненным значением νl, III – расчет по лоренцевскому контуру (χ(Δν)=1); 1 – данные [1], 2 – данные [2], 3 – данные [4,11], 4 – данные [3,4]. Coefficient of absorption by the wings of CO2 lines: a - self-broadening, b - broadening by nitrogen: I - calculation with a theoretical contour, II - calculation with a refined value of νl, III - calculation along the Lorentzian contour (χ (Δν) = 1); 1 - data [1], 2 - data [2], 3 - data [4,11], 4 - data [3,4]. [1] Winters B.H., S. Silverman, W.S. Benedict Line shape in the wing beyond the band head of the 4.3 μ band of CO2 JQSRT V. 4, Issue 4, 1964, Pages 527-537 [2] Burch D.E., Gryvnak D.A., Patty R.R., Bartky Ch.E. Absorption of infrared radiant energy by CO2 and H2O. IV. Shapes of collision -broadened CO2 lines. J.Opt.Soc Amer. 1969, 59, No.3, 267-280 [3] Докучаев А.Б., Тонков М.В. Определение формы крыльев колебательно-вращательных линий полосы двуокиси углерода. Оптика и спектроскопия, 48, вып.4, 738-744 (1980) [4] Sattarov, K., Tonkov, M.V. Infrared absorption in the wing of the v3 vibrational-rotational band of CO2.Oпт. Спектроск. 54 (1983) 944-946.
The absorption by pure CO2 beyond the ν3 bandhead has been measured with a grating spectrometer. Experiments have been made in the 0–60-bar and 291–751-K pressure and temperature ranges. Our room temperature determinations are in good agreement with previous ones and the measured temperature dependence above room temperature is consistent with recent determinations below 300 K. Lorentzian calculations, modified by the introduction of a line shape corrective factor x, are presented. Good agreement between the observed and calculated spectra is obtained when a temperature independent x factor, determined by Cousin et al. at 296 K, is used.
Figure 5. Wavenumber dependence of the pure CO2 normalized absorption coefficient at (a) 291°K; (b) 534°K; (c) 751°K; *, experimental; calculated with the Lorentzian model, Lorentzian model with the χ (296°K) factor of Refs. 4 and 5. 4. R. Le Doucen, C. Cousin, C. Boulet, and A. Henry, "Temperature Dependence of the Absorption in the Region Beyond the 4.3-μm Band Head of CO2.1: Pure CO2 Case," Appl. Opt. 24,897-906 (1985). 5. C. Cousin-Lucasseau, "Absorption I.R. du CO2 dans la fenetre atmospherique autour de 4.2 μm - Determination de la dependance en temperature du coefficient d'absorption.- Influencedes interferences spectrales sur le profil observe," Thesis,Rennes (1987).
We present experimental results on the absorption by CO2-N2 mixtures in the 4.3 μm region. Measurements have been made in the 300–800 K and 0–60 bar ranges; these are in good agreement with previous determinations below 296 K. Frequency- and temperature-dependent factors χ are introduced in order to account for the subLorentzian behavior of the CO2 far line-wings. Their dependences on temperature are deduced from experimental results beyond the 4.3μm band-head in the 193–773 K range for both CO2-CO2 and CO2-N2 and fitted by simple analytical functions. Comparisons are presented between experimental and theoretical spectra on both the low- and high-frequency sides of the band (2100–2600 cm-1 range). It is shown that calculations using χ factors are inaccurate near line-centers at elevated pressures.
Figure 4. Transmission spectra for pure CO2 (in a 4.4 cm long cell) for (a) 291°K and 50 bar and (b) 627°K and 32 bar (), Experimental data; ( ..... ), Eq. (4) with Bi parameters for the given temperature; ( .... ), Eqs. (3) and (4) with the parameters αi, βi, εi of Table 3.
Figure 1. Methane absorption coefficient at 296°K. A. Comparison between experimental and computed spectrum. The computed spectrum is given by the sum of l= 3 (a), l = 4 (b) and double transition (c) components. B. The two components (3, 3) and (3, 4) of the double transition spectrum are shown. The parameters used in the computations are reported in the table (case a).
An attempt is made to describe the observed band shape deformations of a number of CO2 vibro-rotational absorption bands (1.4, 1.6 and 2.0 pm) making use of the ideas on the associative equilibrium in compressed gas. It is shown that the model of monomer-dimer equilibrium permits representation of the IR absorption band shapes within an experimental accuracy up to 5.0 MPa. The comparison of the found dimeric band shape with the calculated synthetic spectra of P- and T-isomers is also discussed.
Figure 3. Comparison of calculated and observed spectra for the 2.0 μm (a) and 1.6 μm (b) CO2 bands. - Measured spectra; o - present work profiles; --- sub-Lorentzian profile.
Spectra of the weakly bound van der Waals complexes H2–N2 and H2–CO have been studied in the mid‐ and far‐infrared regions corresponding to H2 vibrational and pure rotational frequencies, respectively. The experiments were done using a Fourier transform infrared spectrometer to study equilibrium gas samples at low temperature (77 K) with a long absorption path. The spectra of the complexes appear as fine structure located near the peaks of the lines in the collision‐induced spectrum of hydrogen. Compared to earlier studies of these species, the present results have more complete coverage of the various possible transitions, better resolution, and better signal to noise. New structure is reported which corresponds to hindered rotational transitions of the N2 component of an H2–N2 complex and is reminiscent of patterns observed for the N2–Ar and (N2)2 complexes. The relation of these results to far‐infrared measurements of Saturn’s satellite Titan, made by the Voyager spacecraft, is discussed.
Figure 1. Observed spectra of the enhancement by N2 of the collision-induced fundamental band of H2 at 77°K, recorded with a total gas density of about 1.3 amagats and a path length of 154 m. The upper curve (a) is for para-H2 and the lower curve (b) for normal H2. The broad underlying absorption is due to collision-induced absorption in H2-N2 collision pairs. Of interest here is the sharper structure, concentrated near the peak of each collision-induced line, which is due to bound H2-N2 complexes.
Figure 4. Pure CO2 transmission coefficients (52.9 mm long cell) for the pressures of 55 atm. Experimental data, calculated with the Lorentzian model, corrected line mixing model. (a) 294°K (91.4 Am) (b) 493°K (32.8 Am).
Figure 8. The 4-μm continuum region at T = 296°K. (a) Cs vs wavenumber [27] and (b) the absorption coefficient vs wavelength.[37,38].
[27] D. E. Burch and R. L. Alt, AFGL-TR-84-0128, Ford Aerospace and Communications Corporation, Aeronutronic Division (1984). [37] K. O. White, W. K. Watkins, C. W. Bruce, R. E. Meredith and F. G. Smith, Appl. Opr. 17, 2711 (1978) [38] F. S. Mills. Dissertation. The Ohio State University (1975).
Figure 5. Коэффициент поглощения СО2, уширенного N2, для частоты 2387.62 см-1 при разных температурах и давлениях, точки – эксперимент [26], 1988 г.; 1 – расчет с дисперсионным контуром, 2 – расчет [26], 3 – расчет по теории крыльев линий с контуром, полученным в [74], 1982 г. а) Т=296 К, б) – Т=370 К. [26] Hartmann J.M., Rosenmann L., Taine J. Temperature and pressure dependencies of absorption in the narrow R66-R68 window of the 12C16O2 ν3-band, JQSRT 1988, V.40, No.2, p.93-99. [74] Несмелова Л.И., О.Б.Родимова, С.Д.Творогов, Коэффициент поглощения света в крыле полосы 4,3 мкм СО2, в кн.: Спектроскопия атмосферных газов (Наука, Новосибирск, 1982), с.4-16. The absorption coefficient of CO2 broadened by N2 for a frequency of 2387.62 cm-1 at different temperatures and pressures, points - experiment [26], 1988; 1 - calculation with a dispersion contour, 2 - calculation [26], 3 - calculation according to the theory of wings of lines with a contour obtained in [74], 1982 a) T = 296°K, b) - T = 370°K. [ 26] Hartmann J.M., Rosenmann L., Taine J. Temperature and pressure dependencies of absorption in the narrow R66-R68 window of the 12C16O2 ν3-band, JQSRT 1988, V. 40, No.2, p. 93-99. [74] Nesmelova LI, OB Rodimova, SD Tvorogov, Light absorption coefficient in the wing of the 4.3 μm CO2 band, in: Spectroscopy of atmospheric gases (Nauka, Novosibirsk, 1982), p. 4-16.
Thermal emission from the deep atmosphere of Venus can be detected on the nightside around 2.3 μm. The analysis of this radiation requires a reliable knowledge of the absorption in the far wings of the nearby allowed CO2 bands and of the absorption due to collision‐induced bands. We measured absorption coefficients for pure CO2 at pressures varying from 30 to 60 bars in the frequency range 3910–4570 cm−1 at 297.5 K. Values between 1.0 and 1.6 × 10−7 cm−1 amagat−2 are found in the 4100–4500 cm−1 interval where emission from the Venus nightside occurs. The comparison of experimental results with synthetic spectra calculated from a line by line code demonstrates that the Lorentzian line shape strongly overestimates the observed absorption, whereas the use of a χ factor extrapolated from the 3800–4000 cm−1 region does not provide enough opacity.
Figure 2. Binary absorptions coefficients α11(ν) for pure CO2 at 297.5°K in semi logarithmic scale: pluses, experimental results. The upper curve in Figure 2b represents the Lorentzian calculation (i.e., χ=1 in equation (1)). The solid curve in Figure 2a and the lower curve in Figure 2b have been calculated with the Burch et al. [1969] χ-factor. The calculations do not include the very weak contributions from the allowed bands located in the 3900-4700 cm-1 region. Burch, D.E., D.A. Gryvnak, R. R. Patty, and C. E. Bartky, Absorption of infrared radiant energy by CO2 and H2O. IV-Shapes of collision-broadened CO2 lines, J. Opt. Soc. Am., 59, 267-280, 1969.
Relative attenuation contributions at temperature extremes:(a) T = +30.30C, humidity = 14.9 g m- 3; (b) T =-21.4oC, humidity = 0.77 g m- 3
The shape of the far wing of self- and N2-broadened CO2 lines has been investigated in the 2150–2250-cm−1 spectral region, i.e., on the low wavenumber side of the lines of the very intense v3 band of 12C16O2 in a temperature range of atmospheric interest (200–300 K). The experimental results have been compared to calculated values based on the AFGL 1986 compilation. It appears that a symmetrical sub-Lorentzian line shape based on experiments made on the high wavenumber side cannot reproduce experiments. Comparison with experiments made at room temperature shows that the asymmetry of the correcting line shape factor χ strongly increases when decreasing the temperature.
Figure 1. Self-Broadening Experimental and Theoretical Results (Absorption coefficientα, (cm-1)) and δ =[(αobs - αcal)/αobs] in % for different temperatures.
Figure 2. Experimental and Theoretical Values of BCO2+N2 in cm-1 amagat-2 Notes: {δ =[(αobs - αcal)/αobs] in %}. χsym means calculation was made with a symmetrical χ factor deduced from high frequency side experiments; χasym means calculation was made with an optimized asymmetrical Χ factor.
Figure 5. Полный коэффициент поглощения СО2 и его селективная и континуальная составляющие при различных температурах как функции давления, ω=2450 см-1; 1 – κполн, 2 – κконт, 3 – κсел. а) Т=300 К; б) Т=627 К
Figure 3. Спектр пропускания СО2-N2 при Т-291 К, l=4.4 см, точки – эксперимент [4], штриховые – расчет с учетом смешивания линий [4], сплошные кривые – наш расчет: а) 1,1' –ρ=1.62 Амага; 2,2’ –ρ=7.27 Амага; 3,3’ – ρ=17 Амага; б) 1,1' –ρ=29.3 Амага; 2,2’ –ρ=51.5 Амага; 3,3’ –ρ=77.1 Амага [4] J. M. Hartmann. Measurements and calculations of CO2 roomtemperature highpressure spectra in the 4.3 μm region. The Journal of Chemical Physics 90, 2944 (1989); doi: 10.1063/1.455894. http://dx.doi.org/10.1063/1.4558944-2950.
Figure 1. The CO2 absorption coefficient. P=0.70992 atm. a) Lorentzian calculation: line wing theory calculation; b) (κw-κl)/κw: (κexp-κl)/κexp, Ref.2 [2] L.L.Strow and B.M.Gentry, J.Chem.Phys. 84, No.3, pp.1149-1156, 1986
Figure 4. Отклонения рассчитанного коэффициента поглощения от экспериментального для СО2+СО2 и СО2+N2 при различных температурах в микроокнах полосы 4.3 мкм. Коэффициент поглощения κexp [4]. Defined Function = κcalc/κexp. [4] Cousin C., Le Doucen R., Boulet C., Henry A., Robert D. Line coupling in the temperature dependence frequency dependences of absorption in the microwindows of the 4.3 μm CO2 band. JQSRT. 36, No6, 521-538 (1986)
Figure 3. Отклонения рассчитанного коэффициента поглощения от экспериментального для СО2+СО2 и СО2+N2 при различных температурах за кантом полосы 4.3 мкм. Коэффициент поглощения κexp [bib]. [1] Bulanin M.O., Dokuchaev A.B., Tonkov M.V., Filipov N.N. Influence of the line interference on the vibratio-rotation band shapes, JQSRT 31, No.6, 521-543 (1984)
We present a simple model which allows for the computation of the rototranslational band of the collision induced absorption spectra of N2-CH4 pairs at temperatures between 70 and 300 K and at frequencies of up to ~550 cm-1. The agreement with current experimental data has been obtained by adjusting the classical, multipole induced dipolar components by adding semiempirical quantum corrections. We have included a Q6-induction term, never considered before, which we believe is essential for the agreement we have obtained. With the set of temperature-independent parameters derived from fitting the experimental data, our model reproduces recent laboratory measurements within 124-300 K temperature and the 10-600 cm-1 frequency range, within few percent root mean square. The method provides reliable temperature dependence of the absorption coefficient as a function of frequency at temperatures as low as 70 K, which are much lower than those at which laboratory measurements are taken. The work is of significance for modeling the infrared opacity of Titan's atmosphere.
Figure 2. Rototranslational collision-induced spectra of N2-CH4 pairs. Markers denote experimental data by (Dagg et al. 1986) at temperatures of 126°, 149°, 179°, and 212°K. Solid lines show our theoretical results based on fitted parameters μ obtained at this work.
Figure 3. Rototranslational collision-induced spectra of N2-CH4 pairs. Markers denote experimental data by (Birnbaum et al. 1993) at temperatures of 162°K, 195°K, and 297°K. Solid thick lines show our theoretical results, based on fitted parameters m obtained in this work. Separate dipolar contributions are marked with thin lines and denote induction by: dots, N2 Θ, dash-dot-dot, N2 Φ; dashes, CH4 Ω; dash-dot, CH4 Φ; dot-dash-dash, CH4 Q6 ; and long dashes, double transitions.
Figure 4. Comparison of the theoretical estimations based on the current (solid line) and the previous (Courtin 1988) (dashed line) models. The absorption spectra are shown at temperatures of 70°, 120°, and 170°K and frequencies up to 600 cm-1.
Régis Courtin, Pressure-induced absorption coefficients for radiative transfer calculations in Titan's atmosphere, Icarus, Volume 75, Issue 2, August 1988, Pages 245-254, https://doi.org/10.1016/0019-1035(88)90004-8
We present experimental and theoretical studies of medium infrared absorption by pure water vapor. Measurements have been made in the 1900–2600 cm-1 and 3900–4600 cm-1 regions, for temperatures and pressures in the 500–900 K and 0–70 atm ranges, respectively. They are consistent with available data and enable the determination of continuum absorption parameters. It is shown that calculations with line shapes derived from the impact approximation are very inaccurate. Models accounting for the finite durations of collisions and line-mixing through wave-number dependent effective broadening parameters are introduced. The latter have been determined using two different approaches, which are (i) empirical determinations from fits of experimental data and (ii) direct predictions from first principles using a statistical approach. Effective broadening parameters obtained using these two different approaches are in satisfactory agreement for both the temperature and wavenumber dependencies. These data are tested by calculations of continua in various spectral regions and the agreement with measured values is satisfactory. The remaining discrepancies probably result from the influence of the internal structures of the absorption bands considered and thus from the influence of line-mixing. Nevertheless, accurate predictions are obtained in wide temperature and spectral ranges when the total absorption at elevated density is considered. This agreement, which is due to the relatively weak continuum absorption and large contributions of nearby lines, makes the present models suitable for most practical applications involving elevated densities.
Figure 4. Experimental pure H2O transmission spectra for 575 K.
Figure 12. Pure H2O transmissivities in the wing of the ν2 band (2.5 cm-1 resolution): - experimental data; --- calculated from Eqs. (15) and (17) and Table 4. Fig. (12) 575°K and the densities: 10.5 Am; 21.3 Am; and 38.2 Am. Fig. (112) 775°K and the densities: 8.30 Am; 14.2 Am; and 25.6 Am.
Figure 3. A comparison of the monochromatic opacities of CO2 generated with the high-T database for pressure conditions appropriate to the various window regions and alternately room temperature (dashed curve) and a temperature appropriate to the windows (see Fig.1). (a) 2.3 μm window; (b) 1.7 μm window; (c) 1.2 μm window complex.
The shape of the 0003–0000 CO2 band in helium has been investigated at room temperature over an extended range of perturber pressures (0–140 atm). Various and strong deviations from an additive superposition of Lorentzian lines have been observed, due to important line mixing effects enhanced by the specific structure of the R branch in this band.
Figure 6. Observed and calculated absorption coefficient in the region just before the bandhead for conditions: P(CO2)=2.5 atm, P(He)=63.6 atm.
Figure 2. Normalized He-broadened absorption beyond the bandhead. Present experimental results; previous results of Ref. 8; Lorentzian calculation. The bottom gives the ratio of the observed absorption to the Lorentzian prediction. Defined function - B0(exp)/B0(Lor).
[8] Y. I. Baranov, M. O. Bulanin, and M. V. Tonkov, Opt. Spectrosc. 50, 336 (1981).
Figure 5. Absorption in the central gap between P and R branches. Observed absorption coefficient (top) in the region just before the bandhead for various experimental conditions (the bottom gives the ratio of the observed coefficient to a Lorentzian prediction) Defined function - B0(exp)/B0(Lor).
In paper I of this series, important deviations from an additive superposition of Lorentzian profiles were experimentally evidenced in the 00°3–00°0 band of CO2 in He. All the observed deviations are explained by the collision‐induced line mixing effects which schematically transfer intensity from the wing of the band to its central part. The IOS approximation has been found to be insufficient while, the ECS approximation leads to theoretical predictions in good agreement with the experimental data over extended ranges of frequency and perturber pressure. However it must be emphasized that it has been necessary to resort to the method in current use for the determination of the fundamental rates, an ad hoc adjustement starting from the observed linewidths.
Figure 3. Comparison between IOS (infinite order sudden (IOS) approximation ) predictions and experimental absorption in the wing above 6990 cm-1. o, ratio of experimental absorption coefficient to that expected from a sum of Lorentzian lines. V, IOS predictions. (The insert shows the same ratio in the R branch region at high He pressures in order to demonstrate that the transfer of intensity is too important. )
Figure 10. Comparison between experimental and ECS absorption coefficients in the central gap (T=296 K). Experiment; Lorentzian profiles; ECS predictions (based on the optimized Q, rates) [lower curves give the ratio of experimental or ECS results to that expected from a sum of Lorentzian lines].
Figure 3. Comparison between (IOS -DBC) predictions and experimental absorption in the wing above 6990 cm-1. ratio of experimental absorption coefficient to that expected from a sum of Lorentzian lines. In this figure the diagonal elements are no longer obtained from Eq. (3) but deduced from the sum rules (see text). Wkk = -Σl=k dl/dk Wlk IOS/DBC=S’k; Wkk=1.01S’k; Wkk=1.02S’k; Wkk=1.05S’k; Wkk=1.1S’k.
Figure 1. Absorption spectra in the 3ν3 CO2 band perturbed by He at room temperature 297 K. • Experimental values (Ref. 6); — calculated values (Eqs. (22), (24), and (25) and fitted parameters of Table II); --- calculated from the Lorentzian model and HITRAN database (Ref. 8). [6] L. Ozanne, Nguyen-Van-Thanh, C. Brodbeck, J. P. Bouanich, J. M. Hartmann, and C. Boulet, J. Chem. Phys. 102, 7306 (1995). [8] L. S. Rothman, R. R. Gamache, R. H. Tipping, C. P. Rinsland, M. A. H. Smith, D. C. Benner, V. Malathy Devi, J. M. Flaud, C. Camy-Peyret, A. Perrin, A. Goldman, S. T. Massie, L. R. Brown, and R. A. Toth, J. Quant. Spectrosc. Radiat. Trans. 48, 469 (1992).
Figure 2. Absorption spectra in the 3ν3 CO2 band perturbed by He at room temperature 297 K. nCO2=4.61 Am, nHe=409 Am. • Experimental values (Ref. 6); — calculated values (Eqs. (22), (24), and (25) and fitted parameters of Table II); --- calculated from the Lorentzian model and HITRAN database (Ref. 8). [6] L.Ozanne, Nguyen-Van-Thanh, C. Brodbeck, J. P. Bouanich, J. M. Hartmann, and C. Boulet, J. Chem. Phys. 102, 7306 (1995). [8] L.S. Rothman, R. R. Gamache, R. H. Tipping, C. P. Rinsland, M. A. H. Smith, D. C. Benner, V. Malathy Devi, J. M. Flaud, C. Camy-Peyret, A. Perrin, A. Goldman, S. T. Massie, L. R. Brown, and R. A. Toth, J. Quant. Spectrosc. Radiat. Trans. 48, 469 (1992).
We present high density experimental and theoretical results on CO2–He absorption in the v3 and 3v3 infrared bands. Measurements have been made at room temperature for pressures up to 1000 bar in both the central and wing regions of the bands. Computations are based on an impact line-mixing approach in which the relaxation operator is modeled with the energy corrected sudden (ECS) approximation. Comparisons between experimental and calculated results demonstrate the accuracy of the ECS approach when applied to band wings and band centers at moderate densities. On the other hand, small but significant discrepancies appear at very high pressures. They are attributed to a number of reasons which include nonlinear density dependence due to the finite volume of the molecules, neglected contributions of vibration to the relaxation matrix, and incorrect modeling of interbranch mixing.
Figure 5. Absorption coefficients in the central region of the 3ν3 band. experimental; calculated with the: ECS model; Lorentzian model. (a) nCO2= 4.62 Am and nHe=121.2 Am (n˜ He=126.2 Am). (b) nCO2= 4.66 Am and nHe=598.7 Am (n˜ He=720.6 Am).
Figure 7. Absorption coefficients in the central region of the 3ν3 band. o experimental; — calculated with the ECS model corrected for the effective shift Deff . (a) nCO2= 4.61 Am and nHe=364.3 Am (n˜ He=409.4 Am). (b) nCO2=5 4.66 Am and nHe=598.7 Am (n˜ He=720.6 Am).
Figure 9. Absorption coefficients in the central region of the ν3 band. o experimental; calculated with the ECS model corrected for the effective shift Δeff and the interbranch corrective factor -..- bR-P=0.4; —bR-P=0.25; --- bR-P=0.0. (a) nCO2= 2.73*10-5 Am and nHe=603.4 Am (n˜ He=727.2 Am); (b) nCO2= 1.63*10-5 Am and nHe=124.3 Am (n˜ He=129.5 Am); (c) nCO2= 4.25 *10-5 Am and nHe=241.5 Am (n˜ He=261.3 Am).
Figure 4. Portion of observed (CO2)2 +(CO2)3. Note 3:1 spacing of (CO2)3 Q-branches, and sharp appearance of the (CO2)3 rQ0-branch due to l-type doubling (conditions: 1.8% CO2 in He, P0=1 bar).
Figure 3. Portion of observed (CO2)2+(CO2)3 infrared spectrum including central a -type Q-branch of (CO2)2. Note resolved P-branch transitions for (CO2)3. The P-branch transitions are labeled only for subbands where K"≥15 (conditions: 1.8% CO2 in He, P0=1.0 bar).
Figure 3. Portion of observed (CO2)2+(CO2)3 infrared spectrum including central a -type Q-branch of (CO2)2. Note resolved P-branch transitions for (CO2)3. The P-branch transitions are labeled only for subbands where KN≥15 (conditions: 1.8% CO2 in He, P0=1.0 bar).
Figure 10. Unidentified Q-branch like absorptions in the observed (CO2)2+(CO2)3 spectrum. The features labeled QA and QB clearly do not belong to the dimer or cyclic trimer. Moreover, they are within 0.4 cm-1 of the predicted vibrational origin of the trimer asymmetric top isomer from model potentials. Overlap of (CO2)3 transitions with the head of both QA and QB is coincidental (conditions: 1.8% CO2 in He,P0=1 bar)
High resolution infrared spectra of a previously unidentified noncyclic isomer of (CO2)3 have been obtained via direct absorption of a 4.3 μm diode laser in a slit jet supersonic expansion. Two vibrational bands (labeled νI and νIII) are observed, corresponding to the two most infrared active linear combinations of the three constituent CO2 monomer asymmetric stretches: νI is redshifted −5.85 cm−1 from the monomer vibrational origin and is predominately a c‐type band of an asymmetric top, while νIII is blueshifted +3.58 cm−1 and is predominately an a‐type band. Transitions with Ka+Kc=odd (even) in the ground (excited) state are explicitly absent from the spectra due to the zero nuclear spin of CO2; this rigorously establishes that the noncyclic isomer has a C2 symmetry axis. The vibrational shifts and relative intensities of the bands are interpreted via a resonant dipole interaction model between the high‐frequency stretches of the CO2 monomers. Rotational constants are determined by fits of transition frequencies to an asymmetric top Hamiltonian. These results are used to determine vibrationally averaged structural parameters for the complex, which is found to be stacked asymmetric but with C2 symmetry about the b inertial axis. The structural parameters are then used to test several trial CO2–CO2interaction potentials.
Figure 4. Portion of noncyclic (CO2)3 isomer nI band spectrum. The observed spectrum corresponds to a c-type band of an asymmetric top, with a hybrid character of <0.10. The model spectrum (Gaussian width 575 MHz, 0.87 c type, 0.13 a type) correctly reproduces the observed features, including the strong Q branch at 2343.3 cm-1. The gaps in the observed spectrum are due to CO2 monomer transitions ~including hot band and 16O12C17O transitions! that have been excised for visual clarity. Conditions: 2% CO2 seeded in He at 0.8 bar, 20 passes of probe through expansion.
Figure 6. Portion of νIII band of the noncyclic (CO2)3 isomer. Some of the main P-branch transitions of the noncyclic trimer are labeled in the model spectrum, and indicated by arrows in the observed spectrum. Also note the strong Q branch marked with an asterisk in the observed spectrum. This corresponds to none of the predicted features, and probably belongs to a cluster larger than the trimer. Conditions: 0.6% CO2 in He, P0=2.1 bar, 2 passes. Model spectrum: Trot=6.5 K, linewidth is 50 MHz. Hybrid character: 0.78 a type, 0.22 c type.
Figure 3. CO2 line profile adopted here compared to a Lorentz profile, an extrapolation [Fukabori et al., 1986] of the line profile specified by Perrin and Hartmann [ 1989], and the line profile used by Pollack et al. [1993]. M. Fukabori, T. Nakazawa and M. Tanaka, Absorption properties of infrared active gases at high pressures—I. CO2, Journal of Quantitative Spectroscopy and Radiative Transfer, 1986, Volume 36, Issue 3, Pages 265-270, DOI: 10.1016/0022-4073(86)90074-9./ Perrin M.Y. and J.M. Hartmann, Temperature-dependent measurements and modeling of absorption by CO2–N2 mixtures in the far line-wings of the 4.3 μm CO2 band. Journal of Quantitative Spectroscopy and Radiative Transfer, Volume 42, Issue 4, October 1989, Pages 311-317 https://doi.org/10.1016/0022-4073(89)90077-0 Pollack J.B., Dalton J.B., Grinspoon D., Wattson R.B., Freedman R., Crisp D., Allen D.A., Bezard B., deBergh C., Giver L.P., Ma Q., Tipping R.H., Near-infrared light from Venus’ nightside: a spectroscopic analysis, Icarus, 1993, Volume 103, Pages 1-42, DOI: 10.1006/icar.1993.1055.y
The ν2 and 3ν3 bands of CO2 in helium baths at 193 K have studied with a Fourier transform interferometer. The behavior of the band shapes has been explored at moderate densities. The energy corrected sudden (ECS) approximation is used to model the relaxation matrix in order to account for line mixing effects. The basis cross-sections were calculated with the simple power law (P). Computed spectra are in good agreement with the observed ones. Measured broadening coefficients are also comparable with the ones derived from the ECS-P model.
Figure 2. Comparison between experimental, Lorentzian and ECS absorption coefficients in the 3ν3 band region. (a) P[CO2]= 46 Torr., P[He] = 2.5 atm., L = 56 m; (b) P[CO2] = 159 Torr, P[He] = 5 atm., L = 96 m.
We present an experimental study of the self- and N2-broadened H2 O continuum in microwindows within the ν2 fundamental centered at ~1600 cm−1. The continuum is derived from transmission spectra recorded at room temperature with a BOMEM Fourier transform spectrometer at a resolution of ~0.040 cm−1. Although we find general agreement with previous studies, our results suggest that there is significant near-wing super-Lorentzian behavior that produces a highly wave-number-dependent structure in the continuum as it is currently defined.
Figure 5. Self-broadened continuum coefficients from this research, Burch, CKD models, Ma and Tipping, and impact theory.
[5] R. H. Tipping and Q. Ma, Theory of the water vapor continuum and validations, Atmos. Res. 36, 69–94 (1995), and references therein. [8] S. A. Clough, in The water vapor continuum and its role in remote sensing, in Optical Remote Sensing of the Atmosphere, Vol. 2 of 1995 OSA Technical Digest Series Optical Society of America, Washington, D.C., 1995, pp. 76–78. [14] D. E. Burch, Continuum absorption by H2O, Ford Aerontronic Rep. AFGL-TR-81-0300, U.S. Air Force Geophysics Laboratory, Hanscom Air Force Base, Mass., 1982.
The shapes of Q-branches in CO2 spectra in the 580–850 cm−1 region for pure gas and mixtures CO2---He and CO2---Ar have been studied at resolutions up to 0.002cm−1. The branch broadening coefficients were measured in the pressure range from a few torr to 50 atm. Those for 11102-02201 Π-Δ transition are different for pressures below and above 5 atm. At higher pressures the branch broadening coefficients are similar for all Q-branches. The observed Q-branch shape transformation is explained by line mixing effects.
Figure 5. Comparison between experimental and Lorentzian profiles for the Q-branch at 720 cm-1. The insert gives the ratio of experimental absorption coefficient to that expected from a sum of Lorentzian lines.
Figure 6. Comparison between experimental and Lorentzian profiles for the Q-branch at 720 cm-1. The insert gives the ratio of experimental absorption coefficient to that expected from a sum of Lorentzian lines.
Figure 7. Comparison between experimental and Lorentzian profiles for the Q-branch at 618 cm-1. The insert gives the ratio of experimental absorption coefficient to that expected from a sum of Lorentzian lines.
Figure 10. Comparison between experimental and Lorentzian profiles in the wings of the ν2 Q-branch. The insert gives the ratio of experimental absorption coefficient to that expected from a sum of Lorentzian lines. The ratio is only given in the region where the experimental profile is not saturated.
Measurements of pure CO2 absorption in the 2.3-μm region are presented. The 3800–4700-cm−1 range has been investigated at room temperature for pressures in the 10–50-atm range by using long optical paths. Phenomena that contribute to absorption are listed and analyzed, including the contribution of far line wings as well as those of the central region of both allowed and collision-induced absorption bands. The presence of simultaneous transitions is also discussed. Simple and practical approaches are proposed for the modeling of absorption, which include a line-shape correction factor χ that extends to approximately 600 cm−1 from line centers.
Figure 7. Room temperature pure CO2 absorption coefficients for a density of 20 amagats: o, experimental values; solid curve, computed values accounting for allowed and induced transitions with the optimized x factor of Table 2 and the HITEMP database; dashed curve, computed contribution of the far wings of allowed bands centered outside the considered spectral region. The lower plot gives the relative difference between observed and computed spectra.
Measurements of absorption coefficients in the 3v3 band of CO2 at 1.44 μm perturbed by Ar up to 146 bar have been analyzed by using two line-mixing theoretical calculations within the impact approximation. In the first approach, the relaxation operator is treated semi-classically with adiabatic corrections. In the second, the relaxation operator is modelled with the Energy Corrected Sudden (ECS) approximation associated with a power fitting law providing the basic rotational state-to-state rates. Although the line-coupling spectroscopic cross-sections of the two models are significantly different, they both lead to satisfactory agreement with bandshapes at moderate densities (< 100 Amagat). Significant deviations between experimental and calculated spectra appear at higher densities. They are mainly attributed to the probable breakdown of the impact and binary collision approximations and to a number of reasons including an incorrect ECS calculation of the interbranch coupling, the nonlinear density dependence due to the finite volume of the molecules, and the neglect of the unknown imaginary part of the off-diagonal elements in the calculated relaxation matrix.
Figure 10. Absorption coefficients in the central region of the 3ν3 band of CO2 compressed by Ar at high density (n1 = 4.68 Am and n2 = 169.4 Am). • Experimental results; Calculated with the ECS model corrected with the interbranch corrective factor bP-R = 0.20. (a) Only individual lineshifts are taken into account: (b) in addition to the individual lineshifts a corrected lineshift of: -4.5 x 10-3cm-1 Am-1 in the R-branch and -25 x 10-3cm-1 Am-1 in the P-branch are taken into account.
Figure 9. Absorption coefficients in the central region of the 3ν3 band of CO2 compressed by Ar at high density. Experimental; Calculated with the SCA model . (a) n1 =4.78 Am and n2 = 112.3 Am. (b) n1 = 4.68 Am and n2 = 169.4 Am.
Figure 7. Absorption coefficients in the central region of the 3ν3, band of CO2 compressed by Ar at high density. Experimental; Calculated with the Lorentzian model, ECS model. (a) n1 =4.78 Am and n2 = 112.3 Am. (b) n1 = 4.68 Am and n2 = 169.4 Am.
We present an experimental study of the self- and N2-broadened H2O continuum in microwindows within the ν2 fundamental centered at ~1600 cm-1. The continuum is derived from transmission spectra recorded at room temperature with a BOMEM Fourier transform spectrometer at a resolution of ~0.040 cm-1. Although we find general agreement with previous studies, our results suggest that there is significant near-wing super-Lorentzian behavior that produces a highly wave-number-dependent structure in the continuum as it is currently defined.
Figure 7. N2-broadened continuum coefficients from this research, Burch, CKD models, Ma and Tipping, and impact theory. [5] R. H. Tipping and Q. Ma, “Theory of the water vapor continuum and validations,” Atmos. Res. 36, 69–94 (1995), and references therein. [8] S. A. Clough, in “The water vapor continuum and its role in remote sensing,” in Optical Remote Sensing of the Atmosphere, Vol. 2 of 1995 OSA Technical Digest Series ~Optical Society of America, Washington, D.C., 1995, pp. 76–78. [14] D. E. Burch, “Continuum absorption by H2O,” Ford Aerontronic Rep. AFGL-TR-81-0300 (U.S. Air Force Geophysics Laboratory, Hanscom Air Force Base, Mass., 1981).
Figure 5. Measured transmission and computed results obtained accounting for line-mixing by using the models LM-R and LM-S in the 720 cm-1 region for spectra (a) 1 and (b) 84. LM-R and LM-S - two models of accounting for the line-mixing. LM – line-mixing within the first order perturbation theory [Rosenkranz E.W. IEEE Trans. Antennas Propag. AP-23, 498-506, 1975, Smith E.W. J Chem Phys 74. 6658-6673. 1981]. LM-R model, parameters from Rodrigers et al JQSRT 1997 with HITRAN 1996. LM-S model, parameters from Strow L.L. anonymous ftp site, 1994. Table 1. Main characteristics of some of the balloon-borne spectra.
Figure 6. Deviations between measured and computed transmissions in the 720 cm-1 region for spectra (a) 1 and (b) 84 of Figures 5a and 5b with a numerically downgraded resolution of 0.05 cm-1 (FWHM).
We present high-density experimental and theoretical results on CO2---Ar gas-phase absorption in the ν3 and 3ν3 infrared bands. Measurements have been made at room temperature for pressures up to 1000 bar in both the central and wing regions of the bands. A non-linear perturber density dependence of the absorption, clearly shown in the far wing, is attributed to the finite volume of the molecules. Furthermore, experiments show vibrational dephasing and narrowing effects. We have performed line-mixing computations based on the Energy Corrected Sudden approximation (ECS impact model). Significant discrepancies between experimental and calculated spectra appear when pressure increases. We then tested the influence of the finite duration of collision by using interpolations between ECS and quasi-static calculations, and we have evaluated the sensitivity of the band profiles to the interbranch mixing effects. Finally, an effective width is used in order to take other effects into account.
Figure 3. Absorption coefficients in the central region of the 3ν3 band: •, measured; ____, calculated with the ECS model. (a) nCO2=3.26 Am and nAr=283.1 Am (n'Ar=342.4 Am); (b) nCO2=3.26 Am and nAr=545.5 Am (n'Ar=765.6 Am).
Figure 7. ν3 band wing parameters B0CO2-Ar. Experimental results from: Refs.9,23; this work. Calculations with: the ECS impact model; the impact/quasi-static interpolation model; the Lorentzian model. [9] Boissoles J., Menoux V., Le Doucen R., Boulet C., Robert D. Collisionally induced population transfer effects in infrared absorption spectra. II. The wing of the Ar-broadened ν3 band of CO2, J.Chem.Phys. 91, No.4, 2163-2171 (1989) [23] Bulanin M.O., Dokuchaev A.B., Tonkov M.V., Filipov N.N. Influence of the line interference on the vibratio-rotation band shapes, JQSRT 31, No.6, 521-543 (1984).
Table 2. Binary absorption coefficients B0CO2-Ar and density effect parameter cCO2-Ar [see Eq.(11)] in the ν3 band wing. Bibliography
Absorption spectra of gas-phase molecular oxygen and zero air at temperatures of 223 and 283 K have been measured in the laboratory using a coolable multipass-optics gas cell and Fourier transform spectroscopy in the wavelength range 455 to 830 nm (12,000–22,000 cm−1). Net absorption cross sections of the O2A−, B−, and γ-bands at <0.002 nm spectral resolution, and pressures of 100 and 1000 hPa zero air have been determined. Binary absorption cross sections of the collision-induced O4 bands at <0.18 nm spectral resolution and a pressure of 1000 hPa pure oxygen have been determined, with corrections for the O2 γ-band absorption. Calculated integrated absorption intensities and, for the O2A− and B−bands, “effective” Einstein A-coefficients are compared with previous literature values.
Figure 10. Net binary absorption cross sections of the O4 visible bands for 1000 hPa pure oxygen at (a) 223 K and (b) 283 K. Figures 210c and 310d show the cross sections corrected for the O2 -γ-band absorption in the 630.0 nm region for Figures 10a and 110b, respectively. Standard deviations of the cross sections are shown on an expanded scale below each of Figures 10a-310d.
The infrared O–D stretching spectrum of fully deuterated jet-cooled water clusters is reported. Sequential red-shifts in the single donor O–D stretches, which characterize the cooperative effects in the hydrogen bond network, were accurately measured for clusters up to (D2O)8. Detailed comparisons with corresponding data obtained for (H2O)n clusters are presented. Additionally, rotational analyses of two D2O dimer bands are presented. These measurements were made possible by the advent of infrared cavity ringdown laser absorption spectroscopy (IR-CRLAS) using Raman-shifted pulsed dye lasers, which creates many new opportunities for gas phase IR spectroscopy.
Figure 5. (a) D2O cluster spectra as a function of source pressure. Scans are taken at 10, 15, 25, 35, and 45 p.s.i. absolute pressure. Note to two uppermost pressure scans are plotted with the same offset to highlight the ‘‘bulk ice’’ feature that appears for the highest source pressure. (b) D2O cluster spectra obtained with a constant source pressure (35 psia) and increasing concentration of water in the carrier gas.
The Energy Corrected Sudden approach is used in order to deduce collisional parameters and to model infrared quantities in Sigma->Sigma bands of CO2-He and CO2-Ar mixtures at room temperature. Measurements are first used for the determination (from a fit) of the rotational angular momentum relaxation time and of some parameters representative of the imaginary part of the relaxation operator. It is shown that line-broadening data as well as absorption in both the wing and central part of the v3 and 3v3 bands lead to consistent determinations. The model is then used for detailed analysis of line-mixing effects. The influences of pressure, of the band spectral structure, and of the collision partner are studied. Differences between the effects of collisions with He and Ar are pointed out and explained.
Figure 5. Absorption in the wing of the ν3 band: (a): CO2+He, (b) CO2+Ar: measured values; computed results obtained with our values of τJ-1 (0.009 and 0.049 cm-1/Am for He and Ar); have been obtained with twice (He) and half (Ar) these values.
Theoretical results for the far-wing line shapes and corresponding absorption coefficients in the high-frequency wing of the ν3 fundamental band of self-broadened CO2 are presented for a number of temperatures between 218 and 751 K. These first-principles calculations are made assuming binary collisions within the framework of a quasi-static theory with a more accurate interaction potential than in previous calculations. The theoretical results are compared with existing laboratory data and are in good agreement for all the temperatures considered.
Figure 3. Calculated absorption coefficient α(ω) in the 2400–2580 cm-1 spectral region of CO2–CO2 is represented by the solid curve; the experimental data from Ref. 15 are indicated by pluses. (a) T = 296 K, (b) T = 218 K. The absorption calculated assuming a Lorentzian line shape is given by the dashed curve. [15] R. Le Doucen, C. Cousin, C. Boulet, and A. Henry, “Temperature dependence of the absorption in the region beyond the 4.3-μm band head of CO2. I: Pure CO2 case,” Appl. Opt. 24, 897–906 (1985).
Figure 3. Calculated absorption coefficient α(ω) in the 2400–2580 cm-1 spectral region of CO2–CO2 is represented by the solid curve; the experimental data from Ref. 16 are indicated by triangles. (a) T = 291 K, (b) T = 414 K, (c) T = 534 K, (d) T = 627 K, (e) T =751 K. The absorption calculated assuming a Lorentzian line shape . [16] J.-M. Hartmann and M.-Y. Perrin, “Measurements of pure CO2 absorption beyond the ν3 band head at high temperature,” Appl. Opt. 28, 2550–2553 (1989).ibliography
Collision-induced absorption (CIA) by CO2 is measured in the 1100–1600 cm−1range using a Fourier-transform spectrometer with a resolution of 0.5 cm−1. The current measurements, which agree well with previous ones but are more precise, reveal pronounced structures on top of both unresolved Fermi doublet bands consisting of P-, Q-,and R-like branches. Assignment of Q-branches at 1284.75 cm−1and 1387.75 cm−1to (CO2)2 dimers seems highly probable. The nature of other peaks observed in CIA and Raman spectra of the CO2 Fermi doublet region is discussed.
Figure 2. (a) CIA spectrum in the vicinity of lower member of the Fermi doublet. (b) Same as in Fig. 2a for the upper Fermi-coupled band.
New FTIR CO2 high resolution spectra are recorded in argon and nitrogen matrices at 11 K at high dilution. Their evolution with CO2 diffusion at higher temperature is also traced. New observations are discussed in regard to previous works. Even at high dilution (1/10 000), the temperature increase causes appearance of several bands both in argon and nitrogen. Identification of carbon dioxide dimer absorptions is tentatively proposed.
Figure 5. (A) Evolution in ν3 region of a CO2/Ar mixture (1/10 000) kept at 30 K during different time lapses (t).: (a) t=0; (b) t =10 min; (c) t = 20 min; (d) t = 30 min; (e) t = 45 min; (B) comparison of a spectrum of 13CO2/Ar sample (1/10 000) in the ν3 region recorded at 11 K after deposition at 20 K with a spectrum recorded at 11 K after annealing at 30 K. M: monomer; : band due to N2 impurity. D', D" dimer; P pair; X not assigned.
The carbon dioxide dimer spectroscopic patterns are retrieved from the analysis of collision-induced absorption (CIA) spectral bandshape at room temperature. It is shown that the use of the simplified model based on the symmetric-top approximation allows roughly consistent simulation of the observed (CO2)2 dimer spectrum. The rotational constants obtained can be considered as effective thermally averaged constants which characterize dimeric structure, strongly distorted from the ground state. The overall CIA bandshape and the integrated intensity of absorption are broken down into partial contributions from tightly bound and metastable dimers and free-pair states. This approach is shown to be in agreement with a wide range of independent spectroscopic and thermodynamic data.
Figure 3. (b) Comparison of the calculated symmetric-top-like dimeric spectrum (solid line) with the experimental one (dots, see lower trace in Fig. 1) in the vicinity of the lower Fermi-coupled band. The dashed line shows the spectral profile which occurs provided rotational predissociation effect is neglected. Solid and dashed lines in this figure refer to the stick spectrum only and the sticks plus dash wings in the Fig. 3a, respectively. (c) Comparison of the calculated symmetric-top-like dimeric spectrum (solid line) with the experimental one (open circles, see lower trace in Fig. 1) in the vicinity of the upper Fermi coupled band.
This paper aims at examination of the (CO2)2 spectral absorption profile retrieved from the room temperature CO2 collision-induced absorption (CIA) spectrum in the region of the ν1, 2ν2 Fermi doublet. The assignment of (CO2)2 vibrations is discussed, paying due attention to high-resolution CARS observations in the regions of νlow and νup. A variational solution of the anharmonic vibrational problem for the (CO2)2 dimer was obtained using an extended basis set consisting of harmonic and Morse oscillator wave functions. This allowed for a satisfactory description of the vibrational dimer ν1, 2ν2 spectrum from first principles. In particular, we have succeeded in proving that the diffuse absorption band seen in the trough between the major Fermi-coupled CIA bands belongs to combinations of CO2 bending vibrations in a dimer.
Figure 3. Typical CARS spectra taken at high resolution in the regions of νlow and νup (14, 15, 23). Neat CO2 gas was expanded at a stagnation pressure of 1.5 bar through a ω = 90-μm-wide and l=1-mm-long slit nozzle. Both spectra were recorded at a relative distance of z/w = 4 from the nozzle exit.
14. A. A. Vigasin, A. A. Ilyukhin, L. Ramonat, V. V. Smirnov, O. M. Stelmakh, and F. Huisken, Khim. Fiz. 15, 88–95 (1996) [in Russian]. 15. F. Huisken, L. Ramonat, J. Santos, V. V. Smirnov, O. M. Stelmakh, and A. A. Vigasin, J. Mol. Struct. 47, 410–411 (1997). 23. L. Ramonat, Ph.D. Thesis, University of Gottingen, 1997.
Figure 2. Качественная иллюстрация температурной зависимости рассчитанных спектров поглощения СО2 для трех температур за кантом 3ν3 полосы 7000 см-1 (1.4 мкм): а – самоуширение СО2+СО2; б – уширение гелием СО2+Не.
Cavity ring-down spectroscopy has been used to obtain the spectra of transient (O2)2 at 577 and 629 nm, at room temperature and pressures ranging from about 0.25 to 10.0 atm. A selection of these spectra are displayed showing the overlapping monomer and dimer features. Pressure-dependent cross sections have been obtained and Rayleigh extinction has been observed. The derived band parameters are compared to recent results from others. Nous utilisons un spectroscomètre à cavité pour obtenir le spectre du fugace (O2)2 à 577 et 629 nm, à température de la pièce et à des pressions allant de 0,25 à 10,0 atm. Nous présentons une sélection de ces spectres qui montrent un chevauchement des caractéristiques du monomère et du dimère. Nous avons déterminé la dépendance en pression des sections efficaces et observé l’amortissement Rayleigh.
Figure 3. CRDS spectra of (O2)2 a(0; 0)← X(0; 0) (top) and a(0; 1)← X(0; 0) (bottom) at three higher densities.
It is the purpose of this paper to present an infrared absorption spectroscopic study of the state of aggregation of water over a wide range of temperature (25–380°C) and pressure (1–250 bar) along the liquid-gas coexistence curve, and in the supercritical domain. The evolution of the spectral profiles asociated with the internal vibrational modes and with the librational motion has been investigated. In supercritical water, at T=380°C, low pressures (densities) in the range 25–50 bar (0.01–0.05 g.cm−3), only monomeric water is detected. Progressive increase of the pressure (density), from 50 to 250 bar (from 0.05 to 0.4 g.cm−3), shows clearly the appearance of water dimers and trimers. Finally, upon decreasing the temperature (250-25 °C) along the liquid-gas coexistence curve, one observes a continuous evolution of the shape of the infrared spectrum characteristic of the presence of oligomers of increasing size and for temperatures lower than 200°C, the progressive appearance of the hydrogen bond network.
Figure 3. Infrared spectrum in the OH stretching region of supercritical water at T=380°C as function of pressure (density) in the range 200-250 bar (0.1-0.4 g.cm-3). The calculated characteristic frequencies of the OH stretching mode of small water clusters are reported for comparison.
Figure 2. Evolution of the OH anti-symmetric (ν3) and symmetric (Vi) stretching modes, and of the bending mode (ν2) of supercritical water at T=380°C as a function of pressure (density) in the range 25-180 bar (0.01-0.1 g.cm-3). (B) Infrared band intensities (Km.mol-1) associated with the OH stretching and bending modes of small cyclic water clusters as calculated by DFT methods (see text).
The temperature variations in collision-induced absorption (CIA) spectra of carbon dioxide in the region of the Fermi doublet are examined. New FTIR CIA spectra are recorded in the temperature range T=206–296°K. The spectra were subject to decomposition in order to separate true dimer contributions to the CIA profile from the base absorption caused by unbound pairs. The use of statistical physics theory allowed for quite nice reproduction of the observed temperature variations of the normalized dimer intensity.
Figure 2. Temperature variations of the widths (a) and the exponents (b) in the generalized Lorentz formula (1). Empty circles refer to the low-frequency component; filled circles are used for the high-frequency one.
Figure 2. The O2–O2 collision-induced absorption cross-section at room temperature from the UV to the NIR. Assignments and vibrational levels of the bands are also shown.
The far wing shapes of CO2 ν3 band in infrared absorption spectra were calculated accounting for line mixing and finite collision duration effects. The result shows that the impact approximation can be used for the rough estimate of absorption intensity in the far wing region. This approach was used for H2O absorption estimate in the 8–12 μm atmospheric window. According to our calculations the H2O monomer absorption could be responsible for the part of absorption depending on the air pressure. However, the part of absorption depending on H2O pressure squared can hardly be explained in the frame of a model presupposing collisions between monomer molecules.
Figure 3. a) The absorption of H2O vapor due to collisions with N2 molecules. Data calculated with the strong collisions and ABC at Cbranch =0.6 models (1, 2); experimental data (3). b) The absorption of H2O vapor due to collisions with H2O molecules: experimental data (1); data calculated with the strong collisions approximation (2).
Line intensities of the three lowest fundamentals of the 12CH3D Triad are modeled with an RMS of 3.2% using over 2100 observed values retrieved by multispectrum fitting of enriched sample spectra recorded with two Fourier transform spectrometers. The band strengths of the Triad in units of 10−18 cm−1/(molecule cm−2) at 296 K are, respectively, 2.33 for ν6 (E) at 1161 cm−1, 1.75 for ν3 (A1) at 1307 cm−1 and 0.571 for ν5 (E) at 1472 cm−1. The total calculated absorption arising from 12CH3D Triad fundamentals is 4.65×10−18 cm−1/(molecule cm−2) at 296 K. In addition, some 740 intensities of nine hotbands are fitted to 8.1%; most of the hotband measurements belong to 2ν6−ν6 and ν3+ν6−ν3 near 1160 cm−1, 2ν3−ν3 near 1290 cm−1 and ν3+ν6−ν6 near 1304 cm−1. The other observed hotbands are ν5+ν6−ν6, 2ν5−ν5, ν5+ν6−ν5, ν3+ν5−ν3, and ν3+ν5−ν5.
Figure 1. Comparison of observed and simulated profile of a CO2 dimer band.
Using both a difference frequency spectrometer and a Fourier transform spectrometer, we have measured transitions in the 12 20 <-- 01 10 band of carbon dioxide at room temperature and pressures up to 19 atm. The low-pressure spectra were analyzed using a variety of standard spectral profiles, all with an asymmetric component to account for weak line mixing. For this band, we have been able to retrieve experimental line strengths and the broadening and weak mixing parameters. In this paper we also compare the suitability of the energy-corrected sudden model to predict mixing in the two previously measured Q branches 20 00<-- 01 10, the 11 10<--00 00, and the present Q branch of pure CO2, all at room temperature.
Figure 7. Absorption in the 12 20<-01 10 Q branch for various pressures. In each panel there are measured (difference frequency) values, and are measured–calculated deviations obtained with the ECS-EP model, respectively, accounting for and neglecting line mixing.
A new technique is introduced for the acquisition of size-selected neutral cluster spectra on the basis of argon-mediated, population-modulated electron attachment. This method is demonstrated and used to obtain the vibrational spectrum of the neutral water hexamer precursor to the (H2O)6- cluster ion. The mid-infrared spectrum of the neutral species is dominated by four intense features above 3400 cm-1, clearly indicating that significant structural rearrangements occur upon slow electron attachment to form the “magic” hexamer cluster anion. Comparison with previous spectroscopic reports and theoretical predictions indicates that the low-energy “book” isomer is most consistent with the observed band pattern and is suggested to be the species that captures a low-energy electron to form the hexamer anion.
Figure 3. Comparison between vibrational predissociation spectra of (a) (H2O)6-‚ Ar7, reproduced from ref 23, and (b) the neutral (H2O)6 complex, obtained by argon-mediated, population-modulated electron attachment
Temperature (200–300 K) and pressure (70–200 atm) dependent laboratory measurements of infrared transmission by CO2–N2 mixtures have been made. From these experiments the absorption coefficient is reconstructed, over a range of several orders of magnitude, between 600 and 1000 cm-1. The elevated densities used in the experiments (up to 200 atm) magnify the contribution of the wings of the v2 band lines. In order to analyze the spectra, a theoretical model based on the energy corrected sudden approximation is proposed which accounts for line-mixing effects within the impact approximation. This approach uses the model and associated parameters built previously to model Q branches (JQSRT 1999;61:153) but extends it by nowincluding all P, Q, and R lines. No adjustable parameters are used and fundamental properties of the collisional relaxation operator are verified by using a renormalization procedure. Comparisons between measured and calculated spectra confirm that neglecting line-mixing (Lorentzian model) leads to an overestimation of absorption by up to three orders of magnitude in the far wings. On the other hand, the proposed approach leads to satisfactory results both in regions dominated by contributions of local lines and in the wing: measured spectra are correctly modeled over a range where absorption varies by more than four orders of magnitude. The largest discrepancies, which appear about 150 cm-1 from the v2 center, can be due to finite duration of collisions effects or to uncertainties in the experimental determination of very weak absorption.
Figure 6. Normalized absorption in the ν2 region at 296 K for CO2 in 200 atm of N2. Measured values, calculated with the present ECS model and the Lorentzian approach, respectively. (Meas-ECS) relative deviations are given in the lower part of the plot. Defined function = (σObs - σCalc)/σObs
Figure 8. Normalized absorption in the ν2 region at 198 K for CO2 in 105 atm of N2. Measured values, calculated with the present ECS model and the Lorentzian approach, respectively. (Meas-ECS) relative deviations are given in the lower part of the plot. Defined function = (σObs - σCalc)/σObs
Quantum line shape calculations of the rototranslational enhancement spectra of nitrogen-methane gaseous mixtures are reported. The calculations are based on a recent theoretical dipole function for interacting N2 and CH4 molecules, which accounts for the long-range induction mechanisms: multipolar inductions and dispersion force-induced dipoles. Multipolar induction alone was often found to approximate the actual dipole surfaces of pairs of interacting linear molecules reasonably well. However, in the case of the N2-CH4 pair, the absorption spectra calculated with such a dipole function still show a substantial intensity defect at the high frequencies (>250 cm−1) when compared to existing measurements at temperatures from 126 to 297 K, much as was previously reported.
Figure 4. Collision-induced rototranslational absorption band of interacting N2 and CH4 molecules, the binary enhancement spectrum of nitrogen- methane mixtures at the temperatures of 126°, 149°, 179°, and 212°K. Our calculations ͑thin lines͒ are compared with existing measurements ͑heavy͒ ͑Ref. 4͒.
I. R. Dagg, A. Anderson, S. Yan, W. Smith, C. G. Joslin, and L. A. A. Read, Collision-induced absorption in gaseous mixtures of nitrogen and methane, Canadian Journal of Physics, 1986, 64(11): 1467-1474, https://doi.org/10.1139/p86-260.
In spite of decades of extensive studies, the role of water dimers (WD) in the atmospheric radiation budget is still controversial. In order to search for evidence of the dimer in the solar near infrared, high spectral resolution pure water vapour absorption spectra were obtained in laboratory conditions for two different pressures and temperatures in the spectral region 5000–5600 cm−1 (1.785 to 2 µm). The residual was derived as a difference between the measured optical depth and the calculated one for water monomer, using the modified HITRAN database and two different representations of the water vapour continuum: CKD-2.4 (Clough–Kneizys–Davies) and the Ma and Tipping continuum. In both cases the residuals obtained are very similar to those expected from a recent theoretical calculation of the WD absorption. However, the WD band half-width at half maximum (HWHM) and dimerization equilibrium constant, keq, required to provide a best fit to the residual, differ for each case. To be in best agreement with the residual calculated by using the Ma and Tipping continuum, the WD bands HWHM should be ∼28 cm−1, and keq=0.02±0.0035 atm−1and 0.043±0.0055 atm−1for temperatures 342 and 299 K respectively. For the residual calculated using the CKD-2.4 continuum the fitted value of the HWHM is ∼18cm−1, and keq=0.011±0.0025 atm−1(342 K) and 0.018±0.003 atm−1(299 K). It is concluded that a substantial part of the WD absorption is already implicitly included within the CKD-2.4 continuum model. The increase in estimated clear-sky global mean absorption of solar radiation due to WD varies from 0.5% to 2.0%, depending on the set of WD parameters used. On the basis of a comparison of the derived keq values with others in the literature, the higher estimate is favoured.
Figure 4. Corrected residual between measurement and modified-HITRAN calculations with CKD-2.4 included. Bars show the measurement errors and errors of spectral line parameters fitting. (a) 20 hPa, 128 m, 299 K measurement. (b) 98 hPa, 9.7 m, 342 K measurement. To best fit the residual, the dimer spectrum is red-shifted by 12 cm-1 and 5 cm-1 for (a) and (b) respectively.
Figure 5. Corrected residual between measurement and modified-HITRAN calculations with Ma and Tipping continuum included. Bars show the measurement errors and errors of spectral line parameters fitting. (a) 20 hPa, 128 m, 299°K measurement. (b) 98 hPa, 9.7 m, 342°K measurement. To best fit the residual, the dimer spectrum is red-shifted by 9 cm-1 and 5 cm-1 for (a) and (b) respectively.
Figure 15. Differential continuum in terms of optical depth computed from the transmission differences plotted in Figure 14. The estimated uncertainty of the retrieval is indicated by the grey shaded area. At large SZA the retrieved continuum becomes negative due to saturation effects. Both CKD models predict higher continuum absorption than observed, showing good agreement in the band wings.
Figure 17. Differential continuum in terms of optical depth computed from the transmission differences plotted in Figure 16. The estimated uncertainty of the retrieval is indicated by the grey shaded area. While CKD 2.4.1. agrees well with the observation, the continuum computed by MT_CKD_1.0 yields lower continuum than observed. The dimer model predicts an isolated feature near 746 nm, while the retrieval exhibits enhanced absorption around 750 nm.
From the spectrum of water (H216O and H218O) trapped in neon matrix recorded between 50 and 9000 cm−1 29 vibrational transitions from the ground state have been identified for the water dimer. Twenty measured in the mid- and near infrared have been assigned to one-, two- and three-quanta transitions of the intramolecular modes, five in the far infrared to intermolecular modes and four in the mid infrared to binary intra + inter combinations. These assignments are based on 16O/18O isotopic shifts and on the comparison with the spectrum of the acetonitrile:water one to one complex in which the vibrational properties of the water subunit are very close to that of the proton donor molecule in (H2O)2. The comparison of the results in the mid- and far infrared with those obtained in the gas phase shows that the Ne matrix induced perturbations are very small for the intramolecular vibrations and do not exceed 20% for the intermolecular ones. Accordingly this set of data can be used to test the ability for a new version of the Gaussian program to account for the anharmonicity of vibrations and its evolution upon hydrogen-bonding. The results show that the anharmonicity coefficients are generally well reproduced for the intramolecular modes and that the highly anharmonic low frequency intermolecular modes are calculated less than 20% higher than observed in the gas phase.
Figure 1. Temperature effect on some bands of (H218O)2 trapped in Ne (Ne/H218O=800). Spectra recorded at 8°K (lower traces) and 4°K (upper traces) after annealing at 10 K. M: Q(1)-type line of 2ν1 of the monomer. Plot 1. and 301. - (ν1+ν2)/(ν1+ν3) PD; Plot 601. and 401. - ν1/2ν2 PA; Plot 201. and 501. - ν1/2ν1 PD.
Figure 17. Результаты расчетов коэффициента поглощения водяного пара при самоуширении в интервале 3.8-4.2 мкм при различных температурах, полученные в квазистатическом [58] и асимптотическом подходах.
Hydrogen bonding among water molecules is largely responsible for non-ideality in water vapor. Spectroscopic observations in pressurized water vapor using either IR or Raman techniques provide evidence of substantial molecular aggregation up to the critical point and above. The mean size or the molar fractions of the individual aggregates which are formed are, however, very difficult to determine from the analysis of spectra. Present paper aims at the spectral bandshape modeling in sub-critical water vapor making use of available IR absorption spectra in the OH fundamental and overtone. The modeling is supported by quantum-chemical ab initio and anharmonic vibrational calculations for the water dimer, the results of which are compared with previous ab initio and low-temperature laboratory data.
Figure 2. Water vapor absorption in the OH stretching fundamental. Points with error bars stand for experimental data from [9]. Light, dash, and solid traces refer to the monomer, dimer, and resulting spectral profiles, respectively.
The molecular complexes formed between a nitric oxide molecule and the various deuterated isotopomers of the methane molecule have been studied in a supersonic jet expansion. The electronic spectrum arising from the transition corresponding to a 3s←π∗ excitation (Ã Σ+2←X̃ Π2) located on the NO chromophore has been recorded employing resonance-enhanced multiphoton ionization spectroscopy, with each of CH4, CH3D, CH2D2, CHD3, and CD4 as the complexing partner. Rich spectra are obtained, whose appearance changes in a systematic way as the amount of deuteration increases. Unexpectedly, it was possible to record spectra not only in the parent mass channel, but also in various fragment channels; this also led to the identification of some O atom resonances; and their origin is discussed. Discussion is presented of the structure in the spectra, and its possible sources including hindered internal rotation of the methane and NO moieties, overall rotation of the complex, and tunneling. In addition, some guidance has been gleaned from ab initio calculations, and these are discussed in the light of the experimental results.
Figure 3. (1+1) REMPI spectra recorded under three different conditions. In (a) only NO and Ar were present and the spectrum was recorded in the m/z=16 mass channel; in (b) NO, CH4, and Ar were present, and again the spectrum was recorded in the m/z=16 mass channel; in (c) NO, CH4 and Ar were present, with the spectrum being recorded in the m/z=46 mass channel. See text for details.
Figure 4. (1+1) REMPI spectra of the five different isotopomers of NO·methane. Whole scans. Spectra recorded in the parent mass channel, except for the CH3D isotopomer, which was recorded in the CH3D+ masschannel (where a small bleed in of O+ signal is seen from the m/z=16 channel, cf. Fig. 3) —see text.
Figure 5. (1+1) REMPI spectra of the five different isotopomers of NO·methane. Low wavenumber regions. Spectra recorded in the parent mass channel, except for the CH3D isotopomer, which was recorded in the CH3D+ mass channel—see text.
Figure 4. The comparison of spectra in the 0–6000 cm–1 range calculated using the far-wing theory and the empirical model. (a) The entire range, (b) the 1200 cm–1 window (in more detail).
The theoretical approach based on the Energy Corrected Sudden Approximation presented in the previous companion paper is used in order to account for line-mixing effects in infrared bands of CO2. Its performance, which was demonstrated using laboratory spectra is confirmed here by considering atmospheric transmission in the 10–14 μm region. Comparisons are made between forward calculations of atmospheric transmission spectra and values measured using two different solar occultation experiments based on high resolution Fourier transform instruments. The results demonstrate that neglecting line-mixing and using a Voigt model can lead to a very large overestimation of absorption that may extend over more than 300 cm−1 in the wing of the CO2 ν2 band. They also demonstrate the capability of our model to represent accurately the absorption in the entire region for a variety of atmospheric paths. Among positive consequences of the quality of the model, the possibility of retrieving amounts of (heavy) trace gases with weak and broad absorption features is demonstrated.
Existing measurements of the collision-induced rototranslational absorption spectra of gaseous mixtures of methane with helium, hydrogen, or nitrogen are compared to theoretical calculations, based on refined multipole-induced and dispersion force-induced dipole moments of the interacting molecular pairs CH4–He, CH4–H2, and CH4–N2. In each case the measured absorption exceeds the calculations substantially at most frequencies. We present the excess absorption spectra, that is the difference of the measured and the calculated profiles, of these supramolecular CH4–X systems at various gas temperatures. The excess absorption spectra of CH4–X pairs differ significantly for each choice of the collision partner X, but show common features (spectral intensities and shape) at frequencies from roughly 200 to 500 cm−1. These excess spectra seem to defy modeling in terms of ad hoc exchange force–induced dipole components attempted earlier. We suggest that besides the dipole components induced by polarization in the electric molecular multipole fields and their gradients, and by exchange and dispersion forces, other dipole induction mechanisms exist in CH4–X complexes that presumably are related to collisional distortion of the CH4 molecular frame.
Figure 2. Dots: measurements of the rototranslational enhancement absorption spectra of hydrogen-methane gas mixtures at various temperatures (Refs. 4 and 8); solid lines: calculations of the binary rototranslational spectra.
The collision-induced, rototranslational absorption spectrum of compressed methane gas is computed, based purely on the reliably known, leading multipole-induced dipole components of CH4 molecular pairs. In contrast to previous work of the kind no ad hoc empirical corrections of unknown exchange force-induced dipole components are attempted. Not surprisingly, the calculated spectra show a sizeable absorption defect at virtually all frequencies, when compared to existing laboratory measurements. The defect suggests the presence of dipole-induction mechanisms in addition to those due to the leading multipole-induced dipole terms. The excess absorption, the differences between measured and calculated spectra, resembles in certain ways the excess absorption spectra seen at the same frequencies in methane-X gas mixtures, where X stands for helium, hydrogen, or nitrogen, respectively [ Buser and Frommhold J. Chem. Phys. 122 024301 (2005)]. To a large extent, the excess absorption seems to be related to collisional distortions of the tetrahedral frame of the unperturbed CH4 molecule.
Figure 1. Existing measurements [17] at temperatures of 163°, 195°, 243°, and 297°K (dots) are compared to calculated absorption spectra (solid line) in methane.
P. Codastefano, P. Dore, L. Nencini, Temperature dependence of the far-infrared absorption spectrum of gaseous methane, Journal of Quantitative Spectroscopy and Radiative Transfer, Volume 35, Issue 4, April 1986, Pages 255-263, https://doi.org/10.1016/0022-4073(86)90079-8.
Figure 3. Potential energy cuts along the R coordinate for Linear 0-0-0 (i.e.,θ1=θ2=00,φ=00; left panel) and near-equilibrium Slipped Parallel (i.e., θ1=θ2=600, φ=00; right panel) configurations. These cuts are computed using CCSD(T)/aug-cc-pVXZ (X=D,T,Q,CBS), CCSD(T)-F12(a,b)/aug-cc-pVTZ, MP2/aug-cc-pVQZ. We give also the SAPT potential of Bukowski et al.35 for comparison. 1a0=1 bohr=0.529177 A.
35. S.R. Bukowski, J. Sadlej, B. Jeziorski, P. Jankowski, K. Szalewicz, S. A. Kucharski, H. L. Williams, and B. M. Rice, J. Chem. Phys. 110, 3785 (1999). 48. K. A. Peterson, D. E. Woon, and T. H. Dunning, Jr., J. Chem. Phys. 100, 7410 (1994). 49 A. Halkier, T. Helgaker, P. Jørgensen, W. Klopper, and J. Olsen, Chem. Phys. Lett. 302, 437 (1999).
Table 5. Comparison between observation (experimental conditions #5) and simulation. The intensity of the simulated bands were adjusted to match the observed ones
Two collision-induced absorption features of oxygen have been investigated by means of the laser-based cavity ring-down technique at pressures between 0 and 1000 hPa and at temperatures in the range 184–294 K. Peak cross sections, resonance widths and integrated cross sections, as well as spectral profiles, have been determined for the broad O2O2–O2O2 resonances centered at 477 and 577 nm. Results are compared with previous measurements to establish an updated temperature dependence for the cross sections of both resonances, yielding integrated cross sections, that exhibit a minimum near 200 K and that increase in a near-linear fashion in the atmospherically relevant range of 200–300 K. A significant increase in the widths of the resonance profiles upon temperature increase is firmly established. Parameters and temperature-dependent trends for the shape and strengths of the resonances are produced, that can be implemented in cloud retrieval in atmospheric Earth observation.
Figure 4. The absorption profiles of the O2 + O2 feature near 477 nm at 294 K, 230°K and 184°K. Residuals from a comparison to a model function are shown as well. Further details are given in Section 2.5.
Figure 5. The absorption profiles of the O2-O2 feature near 577 nm at 294 K taken from Ref. [2], 268 and 190 K. Residuals from a fit to a model function as discussed in Section 2.5 are shown as well.
Helium droplet technique has been used in order to measure the strength of the infrared absorption in small ammonia and water clusters as a function of size. Hydrogen bonding in ammonia and water dimers causes an enhancement of the intensity of the hydrogen stretching bands by a factor of four and three, respectively. Two types of the hydrogen bonded clusters show different size dependence of the infrared intensity per hydrogen bond. In ammonia (NH3)2 and (NH3)3 it is close to the crystal value. In water clusters, it increases monotonically with cluster size being in tetramers, a factor of two smaller than in the ice. The measured infrared intensity in water clusters is found to be a factor of two to three smaller as compared to the results of numerical calculations.
Figure 1. Depletion spectra of NH3 panel a and H2O panel b single molecules and clusters in He droplets of about 3500 atoms with an average number of about 1.5 and 2 captured molecules per droplet, respectively. Vertical dashed lines give the frequencies of the vibrational band origins of monomers. Sharp rovibrational lines of monomers are assigned Refs. 25 and 26. Additional strong bands are assigned to the absorption of the (NH3)k and (H2O)k clusters, which have been enhanced by hydrogen bonding. Cluster size, k, is indicated by numbers. Spectral peaks a – d in panel b are assigned to the dangling OH bonds of water clusters of different size.
The IR spectra of complexes of water with nitrogen molecules in the range of the symmetric (ν1) and antisymmetric (ν3) bands of H2O have been studied in helium droplets. The infrared intensities of the ν3 and ν1 modes of H2O were found to be larger by factors of 1.3 and 2, respectively, in the N2−H2O complexes. These factors are smaller than those obtained in recent theoretical calculations. The conformation of the N2−H2O complex was estimated. Spectra and IR intensities of the (N2)2−H2O and N2−(H2O)2 complexes were also obtained and their structures are discussed.
Figure 1. Depletion spectra of the N2-H2O complexes in the range of the ν3 band of H2O in He droplets. Pickup pressures: panel (a) PH2O ) 3*10-6 mbar, PN2 = 0; panel (b) PH2O=3 10-6 mbar, PN2 = 9*10-6 mbar. Panel (c) shows spectrum (b) with the contribution of spectrum (a) subtracted. This has been scaled to eliminate H2O linesfrom the spectrum. Smooth curves are Gaussian fits as described in the text. The origin of the ν3 band of H2O in He droplets is marked by the vertical dashed line. Pickup pressure dependences of the intensity of the peaks marked in panel (b) by arrows are shown in Figure 2.
Figure 3. Depletion spectra of the N2-H2O complexes in the range of the î1 band of H2O in He droplets. Pickup pressures: panel (a) P(H2O) = 6*10-6 mbar, P(N2) = 0; panel (b) P(H2O) = 6*10-6 mbar, P(N2) = 9*10-6 mbar; panel (c) P(H2O) = 6*10-6 mbar, P(N2) = 1.8*10-5 mbar. Smooth curves are Gaussian fits as described in the text. The spectra were measured with a collimated laser beam, which gave about a factor of 5 larger effective laser energy flux, as compared with measurements in Figure 1. The origin of the î1 band of H2O in He droplets is marked by the vertical dashed line.
Figure 10. Comparison of experimental data with HITEMP and CDSD (1550 K, 4:3 μm band of CO2). Tashkun SA, Perevalov VI, Bykov AD, Lavrentieva NN, Teffo J-L. Carbon dioxide spectroscopic databank (CDSD), available from hftp://ftp.iao.ru/pub/CDSD-1000i; 2002. Rothman LS, Wattson RB, Gamache RR, Schroeder J, McCann A. HITRAN, HAWKS and HITEMP high temperature database. Proc SPIE 1995;2471:105–11. Bharadwaj SP, Modest MF, Riazzi RJ. Medium resolution transmission measurements of water vapor at high temperature. ASME J. Heat Transfer 2006;121:374–81.
Figure 13. Comparison of experimental data with HITEMP and CDSD (1550 K, 15:0 μm band of CO2). Tashkun SA, Perevalov VI, Bykov AD, Lavrentieva NN, Teffo J-L. Carbon dioxide spectroscopic databank (CDSD), available from hftp://ftp.iao.ru/pub/CDSD-1000i; 2002. Rothman LS, Wattson RB, Gamache RR, Schroeder J, McCann A. HITRAN, HAWKS and HITEMP high temperature database. Proc SPIE 1995;2471:105–11.
We have built new global fits for the ground state potential energy surfaces (PES) of N2–H2 and N2–N2 complexes using ab initio perturbative and supermolecular methods. The analytical expressions used in the four-dimensional fitting procedure require the knowledge of the multipole moments, the static and dynamic multipolar polarizabilities of each monomer, from which long-range electrostatic, induction and dispersion coefficients are evaluated. In agreement with previous work, we have found the most stable conformation of N2–H2 to be linear and that of N2–N2 to have a 45/50° canted parallel shape. The quality of present PESs have been checked by comparing between calculated and experimental second virial coefficients and integral scattering cross-sections, which are found to be in good agreement.
Figure 1. Potential energy as a function of the intermolecular distance R for some selected conformations of the dimer, (θN2a; θH2/N2b; Φ) = (90°; 90°; 0°) for H, (0°, 0°, 0°) for L, (90°, 90°, 90°) for X, (90°, 0°, 0°) for Ta and (0°, 90°, 0°) for Tb, of N2–H2 (left panel) and N2–N2 (right panel).
The present paper considers the density effect on the water vapour absorption profile in the OH stretch range. The density evolution of the water vapour OH fundamental band shape is interpreted in terms of the monomer–dimer equilibrium. In contrast to previous works we adopt experimental values for the frequencies and IR intensities of the dimer vibrations in the vicinity of 3µm. This made it possible to reduce the number of unknown parameters required in the course of our spectral fit.
Figure 2. Typical examples of the spectral fit for the 650K isotherm
Figure 3. Calculated far-IR absorption spectra per molecule for the dimer. The spectrum obtained by Scribano and Leforestier (Ref. 26) (dotted line) is also shown, for the comparison.
Figure 4. Calculated far-IR absorption spectra per molecule for the tetramer and the hexamer
The polarized and depolarized Raman profiles of supercritical CO2 have been measured in the region of the ν2 bending mode (forbidden transition at about 668 cm−1) and for the Fermi dyad (1285 and 1388 cm−1) along the isotherms 307, 309, 313, and 323 K in a reduced density domain 0.04<ρ* = ρ/ρC<2.04 (ρC ∼ 467.6 kg m−3, ρC is the critical density). The spectral features associated with the ν2 mode (degeneracy removal of the mode and Raman intensity activation) are found to be due to the formation of transient complexes. This is supported by the spectral signatures predicted for parallel slipped dimer and trimers (cyclic and noncyclic) from ab initio calculations taking into account the frequency anharmonicity. The band-shape analysis of the Fermi doublet (observed in the spectral range of 1260–1400 cm−1) shows that on the subpicosecond time scale of the Raman spectroscopy, a tagged CO2 molecule probed two kinds of environment in its first shell of neighbors independent of local density enhancement phenomenon. The first one involves interactions of CO2 with surrounding molecules in the first shell whereas the latter is associated with a transient dimer formation. Finally, a broad band observed between the Fermi dyad (at about 1335 cm−1) is assessed from symmetry considerations and from its depolarization ratio as a further evidence of transient complex formation in supercritical CO2.
Figure 5. Influence of the anharmonicity on the ab initio calculated Raman spectrum of noncyclic trimer in the bending region of CO2. As for previous figures, each transition has also been represented by a Lorentzian profile (with an arbitrary full width of 3.0 cm−1).
Intermolecular potentials for the three lowest multiplet states (singlet, triplet and quintet) of the O2(3Σ−g)–O2(3Σ−g) dimer have been investigated in detail by means of high level ab initio calculations. The methods used include MRCI, ACPF, CASPT2, using different active spaces and basis sets. The results for the quintet state are compared with benchmark CCSD(T) calculations. As expected, the former methods do not account accurately for dispersion interactions, although the CASPT2 method performs better than the CI based ones. On the other hand, it is shown that highly correlated methods are necessary to accurately describe the splittings among the multiplet states. We propose to obtain singlet and triplet interaction potentials by combining CCSD(T) quintet potentials and multiconfigurational singlet–quintet and triplet–quintet splittings, respectively. The calculated splittings are quite stable regarding the method employed, except for the well region of the singlet and triplet states within the rectangular configuration, which corresponds to the absolute minima of these multiplet states. Nevertheless, we have been able to assess adequate upper and lower bounds to the interaction potential for this particular region.
Figure. 4 Intermolecular potentials for the the O2–O2 dimer singlet and triplet states as obtained for the four limiting geometries by combining CCSD(T) quintet energies and multireference splittings
Figure 2. Self-continuum absorption of water vapor at P = 2.13 kPa at different temperatures. Data for the 84.1 cm-1 window are not included in the ν2-fit presented by solid curves. The absorbance, A, is expressed as A= log10(1/T), where the maximum transmittance, T, is equal to unity. For convenience, the second scale in dB/km is also presented. For the pathlength used here (23.3 m), the absorbance of A = 1 corresponds to absorption coefficient of 429 dB/km.
Figure 4. Foreign continuum data for two different mixtures of H2O/N2: 1.43/78.5 kPa (triangles) and 0.67/70 kPa (rhombs). For clarity, only the end-temperature data are shown. The curves present the ν2-fit of absorbance data over all available windows.
Figure 7. The spectra of water vapour in 5300 cm-1 (a), and 3700 cm-1 (b) absorption bands, obtained in the experiments of Poberovsky [73, 74] for the cases (1): H2O + N2, and (2): pure H2O (see text); the difference spectrum (2)-(1), attributed by Poberovsky to water clusters; the difference spectrum of Poberovsky, modified in this work; WD according to the model [52]. (a) Experiment [73]: T=530°K, spectral resolution FWHM=15 cm-1, L=4.83 cm (1) and 0.49 cm (2); WD simulation: line strengths and positions from [52], HWHM=28 cm-1 [54], Keq(530°K)=0.0023 atm-1 (Curtiss et al. [59] extrapolation). (b): Experiment [74]: T=503°K, spectral resolution FWHM=10 cm-1; L=0.88 cm (1) and 0.08 cm (2); WD simulation: HWHM=25 cm-1, Keq(503 K)=0.0032 atm-1.
[52] Schofield D.P., Kjaergaard H.G. Calculated OH-stretching and HOH-bending vibrational transitions in the water dimer. Phys. Chem. Chem. Phys., 2003;5:3100–5. [54] Ptashnik I.V., Smith K.M., Shine K.P., Newnham D.A. Laboratory measurements of water vapour continuum absorption in spectral region 5000–5600 cm-1: evidence for water dimers. Q. J. R. Meteorol Soc 2004;130:2391–408. [59] Curtiss L.A, Frurip DJ, Blander M. Studies of molecular association in H2O and D2O vapors by measurement of thermal conductivity. J. Chem. Phys. 1979;71:2703–11 [73] Poberovsky A.V. Problemy fiziki atmosfery. Sbornik Trudov Univ Leningrad 1976;13:81–7 (in Russian). [74] Poberovsky A.V. Diss. kand. fiz-mat nauk (PhD. thesis), Leningrad University, 1976 (in Russian).
A new spectroscopic database for carbon dioxide in the near infrared is presented to support remote sensing of the terrestrial planets (Mars, Venus and the Earth). The compilation contains over 28,500 transitions of 210 bands from 4300 to 7000 cm−1 and involves nine isotopologues: 16O12C16O (626), 16O13C16O (636), 16O12C18O (628), 16O12C17O (627), 16O13C18O (638), 16O13C17O (637), 18O12C18O (828), 17O12C18O (728) and 18O13C18O (838). Calculated line positions, line intensities, Lorentz half-width and pressure-induced shift coefficients for self- and air-broadening are taken from our recent measurements and are presented for the Voigt molecular line shape. The database includes line intensities for 108 bands measured using the McMath–Pierce Fourier transform spectrometer located on Kitt Peak, Arizona. The available broadening parameters (half-widths and pressure-induced shifts) of 16O12C16O are applied to all isotopologues. Broadening coefficients are computed using empirical expressions that have been fitted to the experimental data. There are limited data for the temperature dependence of widths and so no improvement has been made for those parameters. The line intensities included in the catalog vary from 4×10−30 to 1.29×10−21 cm−1/(molecule cm−2) at 296 K. The total integrated intensity for this spectral interval is 5.9559×10−20 cm−1/(molecule cm−2) at 296 K.
Figure 1. Absorption for the nine isotopologues of CO2. The log of the line intensities are plotted vs. cm−1 for the nine species included in the present database. The line intensities vary from 4*10-30 to 1.29*10-21 cm−1/(molecule cm-2) at 296 K. The total integrated strength for this spectral interval is 5.9559*10-20 cm−1/(molecule cm−2) at 296 K.
Carbon dioxide is one of the most important trace gases in the terrestrial atmosphere. The spectral data required in remote sensing are the spectral parameter of each absorption line and a line shape model. This paper describes the absorption properties of CO2 near 2400 cm-1; these properties are of interest to those in the atmospheric temperature sounding field. The shape of the far-wing of N2- and O2-broadened CO2 lines was investigated in the 2200–2500 cm-1 spectral region in a temperature range of atmospheric interest (230–318°K). We focused on the higher rotational quantum number of the R-branch in the v3 band, where the effect of the far-wing is enhanced. The effect of the far-wing has been studied extensively by others, since the CO2 v3 band is known to exhibit sub-Lorentzian behavior. Here, we show the observed spectra along with calculated spectra for five temperatures. We used first-order line-mixing and the x-factor, which accounts for the effect of the far-wing, to create the calculated spectra. Our results provide new knowledge of quantum interference of the spectral line in the v3 band of CO2.
Figure 5. N2-broadened measured spectra (black lines) and spectra calculated with first-order line-mixing coefficients (gray lines) at T = 230, 250, 273, 296, and 318 K.
In this article, we report on a Fourier transform infrared study of absorption bands belonging to small-sized water clusters formed in a continuous slit nozzle expansion of water vapor seeded in argon carrier gas. Clear signatures of free and H-bonded OH vibrations in water aggregates from dimer to pentamer are seen in our spectra. Following an increase in argon backing pressure, the position of the cluster absorption bands varies from those characteristics of isolated water aggregates in the gas phase to those known for clusters trapped in a static argon matrix. These variations can be interpreted in terms of sequential solvation of the water clusters by an increasing number of argon atoms attached to water clusters. Our measured spectra are in good agreement with those obtained previously either for free or Ar coated small-sized water clusters using pulsed slit-jet expansions. Our results are equally in accord with those originating from a variety of tunable laser based techniques using molecular beams or free jets or from the study of water aggregates embedded in rare gas matrices. Distinctions are reported, however, and discussed. Ab initio calculations have made it possible to speculate on the average size of an argon solvation shell around individual clusters as well as on the development of the OH stretch vibrational shifts in mixed (H2O)mArn clusters having different compositions and architectures.
Figure 5. The variation in vibrational frequencies (a) (Ref. 7) and absolute infrared absorption intensities (b) as a function of the number of water molecules in small-sized water clusters. In the lower panel, the measured values from Refs. 30 and 31 are shown by open circles, ab initio calculated intensities from Refs. 32 and 33 are shown by asterisks. 7. F. Huisken, M. Kaloudis, and A. Kulcke, J. Chem. Phys. 104, p.17, 1996. 30. M. N. Slipchenko, K. E. Kuyanov, B. G. Sartakov, and A. F. Vilesov, J. Chem. Phys. 124, p.241101, 2006. 31. S. Kuma, M. N. Slipchenko, K. E. Kuyanov, T. Momose, and A. F. Vilesov, J. Phys. Chem. A 110, p.10046, 2006. 32. S. S. Xantheas and T. H. Dunning, Jr., J. Chem. Phys. 99, p.8774, 1993. 33. M. Losada and S. Leutwyler, J. Chem. Phys. 117, p.2003, 2002.
We report three modifications to recent ab initio, full-dimensional potential energy surfaces (PESs) for the water dimer [X. Huang et al., J. Chem. Phys.128, 034312 (2008)]. The first modification is a refit of ab initio electronic energies to produce an accurate dissociation energy De. The second modification adds replacing the water monomer component of the PES with a spectroscopically accurate one and the third modification produces a hybrid potential that goes smoothly in the asymptotic region to the flexible, Thole-type model potential, version 3 dimer potential (denoted TTM3-F) [G. S. Fanourgakis and S. S. Xantheas, J. Chem. Phys.128, 074506 (2008)]. The rigorous D0 for these PESs, obtained using diffusion Monte Carlo calculations of the dimer zero-point energy, and an accurate zero-point energy of the monomer, range from 12.5 to 13.2 kJ/mol (2.99–3.15 kcal/mol), with the latter being the suggested benchmark value. For TTM3-F D0 equals 16.1 kJ/mol. Vibrational calculations of monomer fundamental energies using the code MULTIMODE are reported for these PESs and the TTM3-F PES and compared to experiment. A classical molecular dynamics simulation of the infrared spectra of the water dimer and deuterated water dimer at 300 K are also reported using the ab initio dipole moment surface reported previously [X. Huang, B. J. Braams, and J. M. Bowman, J. Phys. Chem. A110, 445 (2006)].
Figure 4. (Color online) Expanded portion of the (H2O)2 spectrum shown in Fig. 3. The smooth curve is a smoothed representation of the structured classical spectrum. The vertical lines represent the nearly exact quantum energies (and not intensities) from Ref. 13, described in detail in the text.
Figure 1. The effect of the χ factor on absorption by a single line. Absorption due to a single ‘‘virtual line,’’ with a strength of unity at the line center and a half width of 0.3 cm-1, multiplied by the χ factors shown in Figure 1a. The thick gray line is the unmodified Lorentz line-shape.
Figure 3. Outgoing longwave radiation (OLR) as a function of the surface partial pressure of CO2 (pCO2) for the three parameterizations. Simulations of Earth The details of the simulations are in section 3. Since the atmosphere was not driven to radiativeconvective equilibrium, the OLR does not realistically represent emission from planets’ surfaces but is intended to provide a comparison between the parameterizations. Defined Function = OLRAS - OLRGBKM (W m-2)
Figure 6, Continuum coefficients retrieved from Polar Atmospheric Infrared Radiance Interferometer (PAERI) measurements in (a) the low-wavenumber wing of the ν2 band, (b) the center, and (c) the high-wavenumber wing. The results of previous work (RWW06) [9] and the MTCKD continuum are shown.
[9] P. M. Rowe, V. P. Walden and S. G. Warren, “Measurements of the foreign-broadened continuum of water vapor in the 6.3-μm band at -30oC,” Appl. Opt. 45, 4366-4382 (2006).
The infrared spectrum of the water trimer trapped in solid neon has been identified. Eighteen groups of absorptions between 1600 and 11,000 cm−1 were assigned to one-, two- and three-quanta transitions of the intramolecular modes. Because of the near equivalence of the three molecules and their weak interactions most of these modes correspond to quasi degenerate vibrations involving the bending δ, free OH stretching (OHf) and bonded OH stretching (OHb) of the three subunits at 1608, 3725 and 3525–3473 cm−1, respectively. In the last case the 52 cm−1 splitting is due to the coupling between the OHb oscillators. Calculated anharmonic frequencies correctly agree with these observations and allow to propose a new assignment of the intermolecular modes. Finally combinations of intra + intermolecular transitions were identified and assigned on the basis of calculated anharmonicity coefficients.
Figure 3. Parts of the infrared spectra of H2O/Ne = 1/140 matrix recorded at 3 K before and after annealing at 12.5 K (lower and upper trace, respectively). Sample deposited at 6 K for 6 h at 15 mmol/h. (a–d) Regions 3d, mOHb + d, 2msOHb and 3mOHf, respectively. Note that the mOHb + d region (b) is subdivided in two parts, one relating to ms, the other to ma. The decrease of the signal to noise ratio in (d) after annealing is due to noticeable decrease of the transmission. PA, PD: dimer; Tri: trimer; P: (H2O)n, n > 3.
XVIth Symposium on High Resolution Molecular Spectroscopy (HighRus-2009)
In a previous series of papers, a model for the calculation of CO2-air absorption coefficients taking line-mixing into account and the corresponding database/software package were described and widely tested. In this study, we present an update of this package, based on the 2008 version of HITRAN, the latest currently available. The spectroscopic data for the seven most-abundant isotopologues are taken from HITRAN. When the HITRAN data are not complete up to J″=70, the data files are augmented with spectroscopic parameters from the CDSD-296 database and the high-temperature CDSD-1000 if necessary. Previously missing spectroscopic parameters, the air-induced pressure shifts and CO2 line broadening coefficients with H2O, have been added. The quality of this new database is demonstrated by comparisons of calculated absorptions and measurements using CO2 high-pressure laboratory spectra in the 1.5–2.3 μm region. The influence of the imperfections and inaccuracies of the spectroscopic parameters from the 2000 version of HITRAN is clearly shown as a big improvement of the residuals is observed by using the new database. The very good agreements between calculated and measured absorption coefficients confirm the necessity of the update presented here and further demonstrate the importance of line-mixing effects, especially for the high pressures investigated here. The application of the updated database/software package to atmospheric spectra should result in an increased accuracy in the retrieval of CO2 atmospheric amounts. This opens improved perspectives for the space-borne detection of carbon dioxide sources and sinks.
Figure 3. Measured absorption coefficients (bottom) and associated square density normalized values (top) for pure CO2 at 295.15K and seven total pressures from 15 to 44atm: (b) Dens normalized Abs (10-6 cm-1/atm2) and (a) Abs Coeff (10-2 cm-1).
Figure 5. Absorption coefficients for 1% CO2 in various pressures (from bottom to top: 28.1,47.6, and 76.0 atm) of N2 at 295.15K in the region of the (2v1+v3)2 band. Measured values, calculated values with the old and new version of our database and software, respectively, neglecting line-mixing and taking this process into account.
Figure 8. Absorption coefficients for 10% CO2 in: 29.9 (bottom) and 72.3 atm(top) of N2 at 295.15K in the region of the 3v3 band. Measured values, calculated values with the new version of our database and software, respectively, neglecting line-mixing and taking this process into account.
Figure 9. Absorption coefficients for 1% CO2 in various pressures (from bottom to top: 20.1, 35.6, and 54.0 atm) of N2 at 218.2K in the region of the (2v1+v3)2 band. Black lines are measured values, whereas blue and red lines have been calculated with the old and new version of our database and software, respectively, neglecting line-mixing and taking this process into account.
Water dimers have been assembled in He droplets and studied by infrared laser depletion spectroscopy. All four OH stretching bands of the dimer have been identified in the spectral range 3590–3800 cm−1. Infrared intensities of the bands are also reported. The results are compared with previous measurements and theoretical calculations.
Figure 2. Spectra of the OH-stretching bands of water molecules and clusters in He droplets of about 3500 atoms at spectral resolution of 1 cm-1 upon capture of one and three water molecules on average per droplet in panels a and b, respectively.
From 1984 to 1996 there was rapid progress in the spectroscopy of CO2 dimers and trimers, but since then little work has appeared. Here we report a number of new high-resolution infrared results. For the dimer, a combination band provides the first experimental intermolecular vibrational frequency, and 13C isotopic spectra clarify the role of resonant and non-resonant vibrational shifts. For the cyclic trimer, a new parallel combination band involving an out-of-plane intermolecular mode is observed, and various 13C isotope results are analysed. In addition to these dimer and trimer results, the spectra also contain features which must be due to larger (CO2) N clusters in the range N ≈ 5–15.
Figure 1. Part of the fundamental band of the CO2 dimer for (12C16O2)2, (13C16O2)2, and 12C16O2–13C16O2. An analogous range is covered for each isotopomer, but this is not so obvious since the nuclear spin statistics are different. The asterisks mark lines that are partly or entirely due to the He–CO2 complex. The simulations use parameters from Table 2, an effective rotational temperature of 3 K, and an assumed Gaussian linewidth of 0.0017 cm-1.
Figure 2. Portions of the observed and simulated combination bands of (a) the CO2 dimer and (b) the CO2 trimer.
Figure 3. Three CO2 cluster absorption features that can be assigned as parallel bands of symmetric or nearly symmetric tops. The bands in the upper two panels have similar, but not identical, rotational constants which are appropriate for (CO2)n with n≈5–7. The band in the lower panel corresponds to n≈10.
The interaction-induced dipole moment surface of the van der Waals CH4–N2 complex has been calculated for a broad range of intermolecular separations R and configurations in the approximation of the rigid interacting molecules at the MP2 and CCSD(T) levels of theory using the correlation-consistent aug-cc-pVTZ basis set with the basis set superposition error correction. The simple model to account for the exchange effects in the range of small overlap of the electron shells of interacting molecules and the induction and dispersion interactions for large R has been suggested. This model allows describing the dipole moment of van der Waals complexes in analytical form both for large R, where induction and dispersion have the key role, and for smaller R including whole ranges of their potential wells, where the exchange effects are important. The proposed model was tested on a number of configurations of the CH4–N2 complex and was applied for the analytical description of the dipole moment surface for the family of the most stable configurations of the CH4–N2 complex.
Figure 3. The ab initio calculation of the dipole moment (a) μx and (b) μy for six configurations of the CH4–N2 complex. Solid line—MP2 calculations; circles—CCSDT calculations. The numbers indicate the configurations.
Figure 1. (a.) Total longwave absorption in a pure CO2 gas at 1 bar and 273 K. (b. T=200K, c. T=250K, d. T=300K) Comparison of the CIA absorption shown in (a) with the parameterisation of Kasting et al. (1984) . As can be seen, the difference between the two datasets is extremely large in the 250–500 cm-1 region. Gruszka, M., Borysow, A., 1998. Computer simulation of the far infrared collision induced absorption spectra of gaseous CO2. Mol. Phys. 93, 1007–1016. doi:10.1080/002689798168709. Baranov, Y.I., Lafferty, W.J., Fraser, G.T., 2004. Infrared spectrum of the continuum and dimer absorption in the vicinity of the O2 vibrational fundamental in O2/CO2 mixtures. J. Mol. Spectrosc. 228, 432–440. doi:10.1016/j.jms.2004.04.010 Kasting, J.F., Pollack, J.B., Crisp, D., 1984. Effects of high CO2 levels on surface temperature and atmospheric oxidation state of the early Earth. J. Atmos. Chem. 1, 403–428.
Two quanta transitions involving the vibrational excitation of both proton donor (PD) and proton acceptor (PA) molecules of the water dimer trapped in inert matrices, a particular case of similtaneous transitions, have been identified. They are characterized by weaker intensity and smaller anharmonicity than the usual combinations of PD or PA. In some cases their intensity is strongly enhanced by quasi perfect resonances with PD combinations, as proved by decoupling effects in 18O/16O isotopic mixtures.
Figure 1. Evolution of the spectrum of water trapped in N2 in the regions 5120–5170 (lower frame) and 7190–7150 cm-1 (upper frame) as a function of the 18O/16O isotopic ratio: (a) 0, (b) 1, (c) 3. Lower frame: the isotopic varieties of PA and PD are reported in this order. Upper frame: for natural water the fit of the dimer absorptions in the range 7185–7175 cm-1 has been reported.
Figure 2. Identification of ST’s in Ne matrix at 3K. Lower trace: 16O/18O decoupling effect for the resonant (ν2 + ν3)PDMST(ν2PA + ν3PD) transitions. 18O/16O isotopic ratios of (a) 0, (b) 0.2, (c) 5. The isotopic varieties of PA and PD are reported in this order. Upper trace: ST(ν1PA + ν1PD) observed at 7251.5 cm-1 in absence of resonance. R(0), P(1), Q(1): rovibrational transitions of H2O monomer.
Precise modelling of infrared absorption by carbon dioxide is of primary importance for radiative transfer calculations in CO
-rich atmospheres like those of Venus and Mars. Despite various measurements and theoretical models dedicated to this subject, accurate data at different temperatures and pressures are still lacking in numerous spectral regions. In this work, using two Fourier Transform Spectrometers, we have measured spectra of pure CO
in a large spectral region range, from 750 to 8500 cm
at various densities (3–57 amagat) and temperatures (230–473 K). Comparisons between measured dipolar absorption bands and spectra calculated with the widely used Lorentz line shape show very large discrepancies. This result is expected since the Lorentz approach neglects line-coupling effects due to intermolecular collisions which transfer absorption from the wings to the band center. In order to account for this effect, a theoretical approach based on the impact and Energy Corrected Sudden approximations has been developed. Comparisons of this model with numerous laboratory spectra in a wide range of pressure, temperature and spectral domain show satisfactory agreements for band centers and near wing regions where the impact approximation is valid. However, as expected, due to the breakdown of the impact approximation, the model fails when considering far wing regions. In the absence of precise models accounting for line-mixing
finite collision duration (non impact) effects, empirical approximations are proposed in order to model the far wings.
Figure 3. Normalized absorptioDin the left panel (•red ) and (•blue ) are, respectively, data from Refs. [7] and [10]; data from Refs. [15] and [6]. [6] Burch DE, Gryvnak DA, Patty RR, Bartky CE. Absorption of infrared radiant energy by CO2 and H2O. IV. Shapes of collision-broadened CO2 lines. JOptSocAm1969;59:267–80. [7] Le Doucen R, Cousin C, Boulet C, Henry A.Temperature dependence of the absorption beyond the 4.3 mm band head of CO2. 1:Pure CO2 case. Appl Opt1985;24:897–905. [10] Perrin MY, Hartmann JM. Temperature dependence measurements and modeling of absorption by CO2–N2 mixtures in the far line- wings of the 4.3 mm CO2 band. JQSRT 1989; 42:311–7. [15] Tonkov MV, Filippov NN, Bertsev VV, Bouanich JP, Nguyen Van Thanh, Brodbeck C,et al. Measurements and empirical modeling of pure CO2 absorption in the 2.3 mm region at room temperature: far wings, allowed and collision-induced bands. Appl Opt 1996; 35: 4863–70.
Figure 6. Absorptions of pure CO2 measured and calculated with and without line-mixing effects at 294 K and 51.28 amagat.
Figure 7. Comparison between absorptions of pure CO2 measured and calculated with and without line-mixing effects. Results are obtained, from top to bottom, for (a): T=294K, NCO2 = 35:51 amagat; (b): T=373K, NCO2 = 31:93 amagat; (c): T=473K, NCO2= 23:63 amagat.
Figure 8. Absorption in the high-frequency wing of the ν2 band region of pure CO2 for (a) T=294K, NCO2 =51.28 amagat; (b) T=373K, NCO2 = 31:93 amagat; (c) T=473K, NCO2 = 23.63amagat. There are the experimental values, lines are absorptions calculated with and without taking into account line-mixing effects, respectively.
Figure 9. Normalized absorption in the high-frequency wing of the ν3 band region of pure CO2 at (a) 260K; (b) 296K; (c) 373K; (d) 473K. There are the present measured values, are normalized absorption scalculated using our line-mixing model (see text) and the Lorentz shape, respectively. Values in open circle in (a) are data measured at 258K by [7]. [7] LeDoucen R, Cousin C, Boulet C, Henry A. Temperature dependence of the absorption beyond the 4.3 μm band head of CO2. 1:Pure CO2 case. Appl Opt 1985; 24:897–905.
Figure 10. Pure CO2 normalized absorption coefficients at the ν3 band wing regions for T=295 K (top); T=473 K (bottom). (•) and ( green) are respectively measured and calculated (see text) values. Measured data at room temperature (top) of previous studies are also reported for comparison: ( red) are values of [9] in the left and [7] in the right, (red ) are values from [10]. [7] Le Doucen R, Cousin C, Boulet C, Henry A. Temperature dependence of the absorption beyond the 4.3 mm band head of CO2. 1: Pure CO2 case. Appl Opt 1985;24:897–905. [9] Menoux V, LeDoucen R, Boulet C. Line shape in the low frequency wing of self-broadened CO2 lines. Appl Opt 1987; 26:554–61. [10] Perrin MY, Hartmann JM. Temperature dependence measurements and modelling of absorption by CO2–N 2 mixtures in the far line-wings of the 4.3 mm CO2 band. JQSRT 1989;42:311–7.
Figure 3. Pure CO2 normalized absorption coefficients at the high-frequency wingside of the ν1+ν3 band for (a) T=230K; (b) T=260K; (c) T=295K and (d) T=373K. Measured and calculated values. At room temperature (c), data of Ref. [15] are also reported for comparison. [15] Tonkov M.V., Filippov N.N., Bertsev V.V., Bouanich J.P., Nguyen Van Thanh, Brodbeck C., et al. Measurements and empirical modelling of pure CO2 absorption in the 2.3 mm region at room temperature: far wings, allowed and collision-induced bands. Appl Opt 1996;35: 4863–70.
Recent laboratory observations and advances in theoretical quantum chemistry allow a reappraisal of the fundamental mechanisms that determine the water vapour self-continuum absorption throughout the infrared and millimetre wave spectral regions. By starting from a framework that partitions bimolecular interactions between water molecules into free-pair states, true bound and quasi-bound dimers, we present a critical review of recent observations, continuum models and theoretical predictions. In the near-infrared bands of the water monomer, we propose that spectral features in recent laboratory-derived self-continuum can be well explained as being due to a combination of true bound and quasi-bound dimers, when the spectrum of quasi-bound dimers is approximated as being double the broadened spectrum of the water monomer. Such a representation can explain both the wavenumber variation and the temperature dependence. Recent observations of the self-continuum absorption in the windows between these near-infrared bands indicate that widely used continuum models can underestimate the true strength by around an order of magnitude. An existing far-wing model does not appear able to explain the discrepancy, and although a dimer explanation is possible, currently available observations do not allow a compelling case to be made. In the 8–12 μm window, recent observations indicate that the modern continuum models either do not properly represent the temperature dependence, the wavelength variation, or both. The temperature dependence is suggestive of a transition from the dominance of true bound dimers at lower temperatures to quasi-bound dimers at higher temperatures. In the mid- and far-infrared spectral region, recent theoretical calculations indicate that true bound dimers may explain at least between 20% and 40% of the observed self-continuum. The possibility that quasi-bound dimers could cause an additional contribution of the same size is discussed. Most recent theoretical considerations agree that water dimers are likely to be the dominant contributor to the self-continuum in the mm-wave spectral range.
Figure 10. Smoothed spectra of WD absorption cross-section from Victorova and Zhevakin (V&Zh) [75], Scribano and Leforestier (S&L) [12] and Lee et al. [13] for temperatures around 296°K. The MTCKD [15] continuum model and empirical self-continuum of Burch [45] are also presented for comparison: (A) mm-wave region and (B) mid-to-far infrared spectral regions. The dashed lines in (B) show ratios (right hand axis) of the S&L and Lee et al. calculated spectra to the experimental continuum [45] (above 340 cm-1) and to the MTCKD-1.3 model (below 340cm-1). The equilibrium constant used by S&L: Keq,S&L(~296°K)=0.05 atm-1, is adopted here to represent WD absorption cross-section for all three works. This causes the cross-section curve from Lee et al. to be a factor 1.86 lower than in the original work. The absorption in dB/km from V&Zh and S&L is converted to cm2 molec-1 atm-1 units using formula: 2:3026 x 10-6 [dB/km]* T *Keq,S&L/ (273.15* NA* PWD), where NA=2.687x1019 molec cm-3; T~296°K; PWD is WD partial pressure (atm) assumed in the particular work:8x10-6 in V&Zh and 2.2x10-5 in S&L. The implicit term Keq, S&L/Keq (where Keq is equilibrium constant applied for calculation in particular work) is used in the formula above to bring different calculations to similar conditions.
[12] Scribano Y, Leforestier C. Contribution of water dimers absorption to the millimeter and far infrared atmospheric water continuum. J Chem Phys2007;126:234301. [13] Lee M-S, Baletto F, Kanhere DG, Scandolo S. Far-infrared absorption of water clusters by first-principles molecular dynamics. J Chem Phys 2008; 128:214506. [15] MTCKD Available from web-site: /http:// rtweb.aer.com/ continuum_frame.htmlS. [45] Burch D. Continuum absorption by H2O, Air Force Geophysics Laboratory report,AFGL-TR-81-0300, Hanscom AFB, MA,1981. [75] Viktorova AA, Zhevakin SA. Microradiowave absorption by dimers of atmospheric water vapor. Sov Phys Dokl 1970; 15:852–5.
Figure 8. Comparison of temperature dependencies for pure water vapour continuum absorption in mm-wave and mid-IR spectral ranges. Experimental data for 190 and 239 GHz are from [67] and [68], respectively; CO2 laser absorptions are from [69,70]. Solid lines show the result of calculations using formula (2a) and universal parameters n=1.63 and Do=1700 K. The data for infrared continuum (right axis) are shown in arbitrary units. The Figure is adopted from [63].
Figure 10B. Smoothed spectra of WD absorption cross-section from Scribano and Leforestier (S&L) [12] and Lee et al. [13] for temperatures around 296 K. The MTCKD [15] continuum model and empirical self-continuum of Burch [45] are also presented for comparison in the mid-to-far infrared spectral regions.
[12] Scribano Y., Leforestier C. Contribution of water dimers absorption to the millimeter and far infrared atmospheric water continuum. J. Chem. Phys., 2007;126:234301. [13] Lee M.-S., Baletto F., Kanhere D.G., Scandolo S. Far-infrared absorption of water clusters by first-principles molecular dynamics. J. Chem. Phys., 2008; 128:214506. [15] MTCKD Available from web-site: /http:// rtweb.aer.com/ continuum_frame.htmlS. [45] Burch D. Continuum absorption by H2O, Air Force Geophysics Laboratory report,AFGL-TR-81-0300, Hanscom AFB, MA,1981.
Figure 10. Smoothed spectra of WD absorption cross-section from Victorova and Zhevakin (V&Zh) [75], Scribano and Leforestier (S&L) [12] and Lee et al. [13] for temperatures around 296°K. The MTCKD [15] continuum model and empirical self-continuum of Burch [45] are also presented for comparison: (A) mm-wave region and (B) mid-to-far infrared spectral regions. The dashed lines in (B) show ratios (right hand axis) of the S&L and Lee et al. calculated spectra to the experimental continuum [45] (above 340 cm-1) and to the MTCKD-1.3 model (below 340cm-1). The equilibrium constant used by S&L: Keq,S&L(~296 K)=0.05 atm-1, is adopted here to represent WD absorption cross-section for all three works. This causes the cross-section curve from Lee et al. to be a factor 1.86 lower than in the original work. The absorption in dB/km from V&Zh and S&L is converted to cm2 molec-1 atm-1 units using formula: 2:3026 x 10-6 [dB/km]* T *Keq,S&L/ (273.15* NA* PWD), where NA=2.687x1019 molec cm-3; T~296°K; PWD is WD partial pressure (atm) assumed in the particular work:8x10-6 in V&Zh and 2.2x10-5 in S&L. The implicit term Keq, S&L/Keq (where Keq is equilibrium constant applied for calculation in particular work) is used in the formula above to bring different calculations to similar conditions.
[12] Scribano Y., Leforestier C. Contribution of water dimers absorption to the millimeter and far infrared atmospheric water continuum. J. Chem. Phys., 2007; 126: 234301. [13] Lee M.-S., Baletto F., Kanhere D.G., Scandolo S. Far-infrared absorption of water clusters by first-principles molecular dynamics. J. Chem. Phys., 2008; 128: 214506. [15] MTCKD Available from web-site: /http:// rtweb.aer.com/ continuum_frame.htmlS. [45] Burch D. Continuum absorption by H2O, Air Force Geophysics Laboratory report,AFGL-TR-81-0300, Hanscom AFB, MA,1981. [75] Viktorova A.A., Zhevakin S.A. Microradiowave absorption by dimers of atmospheric water vapor. Sov. Phys. Dokl., 1970; 15:852–855.
Thirteen specific infrared bands in the 2350 cm−1 region are assigned to carbon dioxide clusters, (CO2)N, with N = 6, 7, 9, 10, 11, 12 and 13. The spectra are observed in direct absorption using a tuneable infrared laser to probe a pulsed supersonic jet expansion of a dilute mixture of CO2 in He carrier gas. Assignments are aided by cluster structure calculations made using two reliable CO2 intermolecular potential functions. For (CO2)6, two highly symmetric isomers are observed, one with S6 symmetry (probably the more stable form), and the other with S4 symmetry. (CO2)13 is also symmetric (S6), but the remaining clusters are asymmetric tops with no symmetry elements. The observed rotational constants tend to be slightly (≈2%) smaller than those from the predicted structures. The bands have increasing vibrational blueshifts with increasing cluster size, similar to those predicted by the resonant dipole-dipole interaction model but significantly larger in magnitude.
Figure 3. Observed and simulated spectra showing bands assigned to the S4 isomer of (CO2)6. The perpendicular band in the lower panel is overlapped by bands of (CO2)2 and of the cyclic (brown) and noncyclic (blue) isomers of (CO2)3 as shown, while the Q- and R-branches of the parallel band in the upper panel are relatively clear. Observed and simulated spectra showing bands assigned to the S4 isomer of (CO2)6. The perpendicular band in the lower panel is overlapped by bands of (CO2)2 and of the cyclic (brown) and noncyclic (blue) isomers of (CO2)3 as shown, while the Q- and R-branches of the parallel band in the upper panel are relatively clear.
The infrared spectrum of the water dimer trapped in solid neon has been recorded up to the visible by improving significantly the experimental technique used in a previous paper [Y. Bouteiller, J.P. Perchard, Chem. Phys. 305 (2004) 1]. A total of 22 intramolecular transitions of the proton donor (PD) and 23 of the proton acceptor (PA) are now identified and assigned on the basis of 16O/18O isotopic shifts and of realistic anharmonicity corrections. From an ab initio determination of the potential energy a perturbation-resonance treatment has been carried out for each polyad Pn, n = 2–8. Finally combinations of intra + intermolecular transitions were identified and assigned on the basis of calculated anharmonicity coefficients.
Figure 1. Parts of the infrared spectrum of a H2O/Ne = 1/150 matrix recorded at 3 K at the resolution of 1 cm-1 in the P6 and P8 domains. Sample deposited at 5.8 K for 6 h at a rate of 7 mmol/h. Tri: water trimer. (b) Spectra recorded before (upper trace) and after annealing at 12 K (lower trace).
Figure 3. Spectrum of H2O trapped in Ne in the P5 region. Same conditions of deposition as for Fig. 1. (a) Spectra recorded at the resolution of 0.5 cm-1. (b) Spectra recorded before (upper trace) and after annealing at 12 K (lower trace).
We employ recent flexible ab initio potential energy and dipole surfaces [Y. Wang, X. Huang, B. C. Shepler, B. J. Braams, and J. M. Bowman, J. Chem. Phys. 134, 094509 (2011)10.1063/1.3554905] to the calculation of IR spectra of the intramolecular modes of water clusters. We use a quantum approach that begins with a partitioned normal-mode analysis of perturbed monomers, and then obtains solutions of the corresponding Schrödinger equations for the fully coupled intramolecular modes of each perturbed monomer. For water clusters, these modes are the two stretches and the bend. This approach is tested against benchmark calculations for the water dimer and trimer and then applied to the water clusters (H2O)n for n = 6–10 and n = 20. Comparisons of the spectra are made with previous ab initio harmonic and empirical potential calculations and available experiments.
Figure 4. Local-monomer IR spectra in the OH-stretch region of the indicated isomers of the hexamer obtained with the local-monomer model.
Figure 6. Local-monomer IR spectra of intramolecular modes of the lowest energy isomer of the indicated cluster.
The origin of the line shape of the O−H stretch vibrational spectrum is analyzed for supercritical water in the low- and medium-density region by using classical molecular dynamics simulation for the flexible point-charge model, SPC/Fw. The spectrum calculated for the water model is in good agreement with the experimental one in the low-density region. The spectral origins in the low-density region of 0.01–0.04 g cm−3 are assigned to a sharp peak due to the bond oscillation along the O−H vector and two broad bands due to the rotational coupling, by taking an isolated single molecule as a reference in the low-density limit. The bands due to the rotational coupling reduce in intensity with increasing density as the rotations are more hindered by the hydrogen-bonding interactions, and their intensities increase with increasing temperature due to the accelerated rotational motion. The O−H stretch oscillation in the time correlation function attenuates in a timescale comparable with the lifetime of the hydrogen bonds, and the spectra conditioned by the number of hydrogen bonds are dominantly controlled by the local solvation structure.
Figure 1. Comparison of the spectrum of the dipole time correlation function ◦calculated at 400◦C with the experimental IR spectrum. The experimental data in (a) by Tassaing et al. (Ref. 6) are taken at 380◦C with 0.009 g cm−3 and those by Vigasin et al. (Ref. 7) at 377◦C with 0.0105 g cm−3. In (b), the data by Tassaing et al. are at 380◦C with 0.04 g cm−3 and those by Vigasin et al. are at 377◦C with 0.0405 g cm−3. The calculated spectrum is shifted by 50 cm−1 to the higher frequency for comparison of the line shape to the experimental one.
A recently introduced bond–bond formulation of the intermolecular interaction has been extended to six-atom systems to the end of assembling a new potential energy surface (PES) and has been incorporated into a grid empowered simulator able to handle the modeling of the CO2 + CO2 processes. The proposed PES is full dimensional and accounts for the dependence of the intermolecular interaction on some basic physical properties of the colliding partners, including modulations induced by the monomer deformation. The used analytical formulation of the interaction involves a limited number of parameters, each having a clear physical meaning. Guess values for these parameters can also be obtained from analytical correlation formulae. Such estimates can then be fine tuned by exploiting experimental and theoretical information. The resulting PES well describes stretched and bent asymptotic CO2 monomers as well as the CO2–CO2 interaction in the most and less stable configurations. On this potential massive quasiclassical elastic and inelastic detailed scattering trajectories have been integrated, by exploiting the innovative computational technologies of the grid. The efficiency of the approach used and the reliability of the estimates of the dynamical properties obtained in this way is such that we can now plan a systematic evaluation of the state specific rate coefficient matrix elements needed for space craft reentry modeling. Here, we present probabilities and cross sections useful to rationalize some typical mechanisms characterizing the vibrational transitions of the CO2 + CO2 system on the flexible monomer proposed PES. On such PES, the key dynamical outcomes are: (a) there is a strong energy interchange between symmetric stretching of the reactants and bending of the products (and viceversa) while asymmetric stretching is strongly adiabatic (b) reactant energy is more efficiently allocated (with respect to the rigid monomers PES) as product vibration when reactant stretching modes are excited while the contrary is true when the reactant bending mode is excited.
Figure 4. Comparison of the interaction energies for selected dimer geometries of a pair of rigid monomers (black lines) and of a rigid plus a stretched (both CAO bonds of monomer a have been symmetrically stretched of 10% with respect to equilibrium) monomer (red dashed lines). The Ta configuration refers to that with the stretched monomer a and the unstretched monomer b perpendicular to it and lying on the line connencting the two molecular centers of mass. Bond-bond results are plotted in the lower panel while those corresponding to the ab initio calculations are plotted in the upper panel.
Figure 5. Comparison of the interaction energies for selected dimer geometries of a pair of rigid monomers (black lines) and of a rigid plus a stretched (the two C-O bonds of monomer a have been respectively stretched and shrunk of 10% with respect to equilibrium) monomer (red dashed lines). In particular the stretched C-O bond is the one lying on the intermolecular separation R and pointing closer to the monomer b in the L configuration. The Ta configuration refers to that with the stretched monomer a and the unstretched monomer b perpendicular to it and lying on the line connecting the two molecular CM. Bond-bond results are plotted in the lower panel while those corresponding to present ab initio calculations are in the upper panel.
An isotopic-independent, highly accurate potential energy surface (PES) has been determined for CO2 by refining a purely ab initio PES with selected, purely experimentally determined rovibrational energy levels. The purely ab initio PES is denoted Ames-0, while the refined PES is denoted Ames-1. Detailed tests are performed to demonstrate the spectroscopic accuracy of the Ames-1 PES. It is shown that Ames-1 yields σ rms (root-mean-squares error) = 0.0156 cm−1 for 6873 J = 0–117 12C16O2 experimental energy levels, even though less than 500 12C16O2 energy levels were included in the refinement procedure. It is also demonstrated that, without any additional refinement, Ames-1 yields very good agreement for isotopologues. Specifically, for the 12C16O2 and 13C16O2 isotopologues, spectroscopic constants G v computed from Ames-1 are within ±0.01 and 0.02 cm−1 of reliable experimentally derived values, while for the 16O12C18O, 16O12C17O, 16O13C18O, 16O13C17O, 12C18O2, 17O12C18O, 12C17O2, 13C18O2, 13C17O2, 17O13C18O, and 14C16O2 isotopologues, the differences are between ±0.10 and 0.15 cm−1. To our knowledge, this is the first time a polyatomic PES has been refined using such high J values, and this has led to new challenges in the refinement procedure. An initial high quality, purely ab initiodipole moment surface (DMS) is constructed and used to generate a 296 K line list. For most bands, experimental IR intensities are well reproduced for 12C16O2 using Ames-1 and the DMS. For more than 80% of the bands, the experimental intensities are reproduced with σ rms(ΔI) < 20% or σ rms(ΔI/δ obs) < 5. A few exceptions are analyzed and discussed. Directions for future improvements are discussed, though it is concluded that the current Ames-1 and the DMS should be useful in analyzing and assigning high-resolution laboratory or astronomical spectra.
Figure 3. Line list comparison at 296 K, Ames-296 K (red) vs. HITRAN-2008 (black). Convolution FWHM = 10 cm−1. Full scale comparison is shown in (a), and amplified, detailed comparisons are given for 0–3000 cm−1 in (b), and 8000–13 000 cm−1 in (c).
Figure 1. Absorption coefficient of pure CO2 at 296 K due to the single 3ν3 band calculated with the present rCMDS model and the ECS approach of Ref. 16 for the densities of (a) 22.65 Am and (b) 51.28 Am. 16. H. Tran, C. Boulet, S. Stefani, M. Snels, and G. Piccioni, J. Quant. Spectrosc. Radiat. Transf. 112, 925 (2011).
Figure 2. Absorption coefficients of pure CO2 in the region of the 3ν3 band at 294 K for densities of (a) 22.7 Am, (b) 35.5 Am, and (c) 51.3 Am. The blue circles are measured values [16] while the lines are calculated results obtained with the rCMDS model (red) and neglecting line-mixing (black). The olive curve in (c) has been obtained from rCMDS after the introduction of spectral shifts of opposite signs in the P and R branches (see text). 16. H. Tran, C. Boulet, S. Stefani, M. Snels, and G. Piccioni, J. Quant. Spectrosc. Radiat. Transf. 112, 925 (2011).
Figure 3. Absorption coefficients of pure CO2 in the region of the 2ν1 + 2ν2 + ν3 band at 294 K for densities of (a) 20.6 Am, (b) 33.0 Am, and (c) 56.7 Am. Measured values [16]. Calculated results obtained with the rCMDS model and neglecting line-mixing. Calculated results obtained from rCMDS after the introduction of spectral shifts of opposite signs in the P and R branches (see text). 16. H. Tran, C. Boulet, S. Stefani, M. Snels, and G. Piccioni, J. Quant. Spectrosc. Radiat. Transf. 112, 925 (2011).
Water dimers (H2O)2 are believed to affect Earth’s radiation balance and climate, homogeneous condensation, and atmospheric chemistry. Moreover, the pairwise interaction which binds the dimer appears to be of paramount importance for expounding a complete molecular description of the liquid and solid phases of water. However, there have been no secure, direct observations of water dimers at environmentally relevant temperatures despite decades of studies. We report the first unambiguous observation of the dimer spectrum recorded in equilibrium water vapor at room temperature.
Figure 2. The observed dimer spectrum (a). The ab intio calculated [21] dimer spectrum [(b) upper trace]. The water vapor continuum absorption measured in moist nitrogen [28] at atmospheric pressure (d). The calculated collision-induced absorption [32] [(c) lower trace], the absorption is shown multiplied by factor of 1000 to make it visible). All spectra correspond to water vapor at 13 Torr and 296 K.
[21] Y. Scribano and C. Leforestier, J. Chem. Phys. 126, 234301 (2007). [28] M. A. Koshelev, E. A. Serov, V. V. Parshin, and M. Yu. Tretyakov, J. Quant. Spectrosc. Radiat. Transfer 112, 2704 (2011). [32] C. Leforestier, R. H. Tipping, and Q. Ma, J. Chem. Phys. 132, 164302 (2010).
Extensive experimental studies of room-temperature carbon dioxide absorption coefficients are reported for a wide range of wavenumbers and pressures requested in atmospheric spectra modelling. The quality of measurements is optimised by the use of two complementary setups with long- and short-path optical cells for low and high gas densities, respectively. The recorded spectra provide a representative picture of band-shape evolutions from gaseous to nearly liquid phases of CO2 and enable a theoretical analysis of line-mixing effects. Various kinds of vibrational bands (Σ←Σ, Π←Σ as well as Π←Π transitions) are modelled using a specific, non-Markovian in the general case, approach of Energy-Corrected Sudden type which is based on the symmetric relaxation matrix and, in contrast to the standard ECS model used for infrared absorption calculations, ensures automatically the fundamental relations of detailed balance and double-sided sum rules. Moreover, this method properly accounts for the vibrational angular momenta of the initial and final molecular states and allows including Coriolis resonances via the usual Herman-Wallis factors in the dipole transition moments. With a set of ECS parameters previously obtained for isotropic and anisotropic Raman spectra modelling, completely neglected imaginary part of the relaxation operator and a simple change in the tensorial rank to get the dipole absorption case in the working formulae, the computed spectra reproduce quite correctly the vibrotational band shapes up to 20 amagat without any additional parameter. An empirical correction factor tentatively introduced to account globally for the Coriolis effects on the relaxation matrix leads to better matches with high-density band shapes but its role merits further studies with an accurately modelled imaginary part of the relaxation matrix.
Figure 4. Dependence of the (ν1 + ν2)I (a) and 3ν1 + ν3 (b) band-shapes on the gas density.
Figure 7. Comparison of calculated and experimental absorptions for the 3ν3 band
Figure 9. Comparison between calculated and experimental absorptions for the (ν1 + ν2)I band. T=300 K.
The collisions between two oxygen molecules give rise to O4 absorption in the Earth atmosphere. O4 absorption is relevant to atmospheric transmission and Earth’s radiation budget. O4 is further used as a reference gas in Differential Optical Absorption Spectroscopy (DOAS) applications to infer properties of clouds and aerosols. The O4 absorption cross section spectrum of bands centered at 343, 360, 380, 446, 477, 532, 577 and 630 nm is investigated in dry air and oxygen as a function of temperature (203–295°K), and at 820 mbar pressure. We characterize the temperature dependent O4 line shape and provide high precision O4 absorption cross section reference spectra that are suitable for atmospheric O4 measurements. The peak absorption cross-section is found to increase at lower temperatures due to a corresponding narrowing of the spectral band width, while the integrated cross-section remains constant (within o3%, the uncertainty of our measurements). The enthalpy of formation is determined to be ΔH250 = -0.12±0.12 kJ mol-1, which is essentially zero, and supports previous assignments of O4 as collision induced absorption (CIA). At 203°K, van der Waals complexes (O2-dimer) contribute less than 0.14% to the O4 absorption in air. We conclude that O2-dimer is not observable in the Earth atmosphere, and as a consequence the atmospheric O4 distribution is for all practical means and purposes independent of temperature, and can be predicted with an accuracy of better than 10-3 from knowledge of the oxygen concentration profile.
Figure 4. Comparison of peak and integrated cross-sections to available literature values at four wavelengths
Измерено поглощение излучения углекислым газом в области 8000 см–1, проведен расчет коэффициентов поглощения с применением асимптотической теории крыльев линий на основе подгонки к экспериментальным данным. Результаты расчета хорошо согласуются с экспериментом. Согласно теории крыльев линий поглощение в крыльях полос обусловлено крыльями сильных линий близлежащей полосы. В рамках этих представлений экспериментальные и расчетные данные о коэффициенте поглощения СО2 в крыльях двух полос в области 8000 см–1 могут являться источником сведений о форме контура спектральных линий при смещенных частотах, отвечающих нескольким десяткам полуширин. Полученные результаты подтверждают гипотезу о том, что параметры контуров в крыльях полос, отвечающих переходам с одним и тем же начальным состоянием, оказываются близкими. Выражение для контура спектральных линий при больших смещенных частотах может быть полезно для расчетов пропускания излучения в окнах прозрачности атмосфер, содержащих СО2.
Figure 8. Коэффициент поглощенияСО2 при самоуширении в крыльях полос в области 8300 см-1 (а) и 8200 см-1 (б) ; (а,б) экспериментальные данные, расчет с HITRAN-2004, (б) расчет с HITRAN-2012; Т=290K, РСО2=612 мбар
Figure 2. Collision-induced absorption spectrum (in cm−1 /amagat2 ) of N2 by N2 (a) at 149 K, (b) at 228 K, and (c) at 296 K, as obtained from measurements12, 13 (circles) and from the CMDS calculations using the present ab initio dipole moment and the intermolecular potential of Ref. 16 (continuous lines). The error bars correspond to ± 10%.
Figure 3. Self- and foreign-continuum cross-section (upper and lower panels respectively) as retrieved from laboratory measurements in Ptashnik et al. (2011, 2012) (CAVIAR), along with their associated uncertainties, compared to the far-wing model of Tipping and Ma (1995) and the MTCKD-2.5 continuum model.
Ptashnik I.V., McPheat R.A., Shine K.P., Smith K.M., Williams R.G. (2011b) Water vapor continuum absorption in near-infrared windows derived from laboratory measurements. J Geophys Res 116:D16305 Ptashnik I.V., McPheat R.A., Shine K.P., Smith K.M., Williams R.G. (2012) Water vapour foreign continuum absorption in near-infrared windows from laboratory measurements. Phil Trans Roy Soc A (to appear). doi:10.1098/rsta.2011.0218 Tipping R.H., Ma Q. (1995) Theory of the water vapor continuum and validations. Atmos Res 36:69–94
Cw-CRDS spectra of water-rare gas supersonic expansions were recorded between 7229 and 7262 cm−1. The effective absorption pathlength was about 1 km in jet-cooled gas and the resolution about 1 × 10−4 cm−1. Many well-resolved structures are observed that could be assigned from experimental evidence to H2O-Ar/Kr bands. Eight broader unresolved features are more specifically reported and assigned to small H2O multimers, in good agreement and refining previous observations by Nizkorodov et al. [J. Chem. Phys.122, 194316 (2005)]. Among these, the band at 7256.5 cm−1 is shown to be a Q branch of the water dimer with accompanying R and very weak P lines. The band is assigned to a K a = 0 ← 1 transition and rotationally analyzed, leading to a restricted set of upper state rotational constants. The upper state lifetime (60 ± 3 ps) is extracted from the linewidths.
Figure 2. Focus on the rotational structure of the (H2O)2 band observed around 7256.5 cm−1 in the cw-CRDS spectra of a supersonic expansion of H2O in Ne, Ar, and Kr carrier gases. Monomer lines and H2O-rare gas lines are also observed. The simulated structure is presented at the bottom.
We present an extensive study of the four-dimensional potential energy surface (4D-PES) of the carbon dioxide dimer, (CO2)2. This PES is developed over the set of intermolecular coordinates. The electronic computations are carried out at the explicitly correlated coupled cluster method with single, double, and perturbative triple excitations [CCSD(T)-F12] level of theory in connection with the augmented correlation-consistent aug-cc-pVTZ basis set. An analytic representation of the 4D-PES is derived. Our extensive calculations confirm that “Slipped Parallel” is the most stable form and that the T-shaped structure corresponds to a transition state. Later on, this PES is employed for the calculations of the vibrational energy levels of the dimer. Moreover, the temperature dependence of the dimer second virial coefficient and of the first spectral moment of rototranslational collision-induced absorption spectrum is derived. For both quantities, a good agreement is found between our values and the experimental data for a wide range of temperatures. This attests to the high quality of our PES. Generally, our PES and results can be used for modeling CO2 supercritical fluidity and examination of its role in planetary atmospheres. It can be also incorporated into dynamical computations of CO2 capture and sequestration. This allows deep understanding, at the microscopic level, of these processes.
FIGURE 4. Potential energy cuts along the R coordinate for Linear 0-0-0 (left panel) and Slipped Parallel 60-60-60 (right panel) configurations. These cuts are computed using CCSD(T)-F12b/aug-cc-pVTZ. We give also the individual contribution of each multipolar expansion term. See text for more details. 1a0 = 1 bohr = 0.529177 Å.
Figure 2. CH4–CO interaction energies for different basis sets and methods as a function of the intermolecular distance for selected angular orientations.
A line-by-line calculation of the continuum absorption coefficient with the wing line contour describing the absorption in different windows outside of the water vapor bands in the 4000–8000 cm-1 spectral region is presented. The continuum absorption calculated with the line wing contour characterizing the absorption in the 5800–6400 cm-1 window is shown to be close to the total absorption. Hence it follows that the absorption in the 5000–5500 cm-1 and 6900–7700 cm-1 bands is almost entirely due to metastable dimers and free complexes.
Table 6. Experimental Н2О self-continuum (T =296 K) versus predicted absorption spectra of bound and metastable dimers1 and present calculations for the 5200–5500 cm-1 band. The black dots are the experimental data, the triangles are the stable dimer absorption coefficients, and the gray curves represent the calculated results using the κLorχ line shape. The frequency step is 50 cm-1.
Non-intrusive spectroscopic probing of weakly bound van der Waals complexes forming in gaseous carbon dioxide is generally performed at low pressures, for instance in supersonic jets, where the low temperature favors dimers, or in few-atmosphere samples, where the signature of dimers varying as the squared gas density is entangled with the dominating collision-induced absorption. We report experimental and theoretical results on CO2 dimers at very high pressures approaching the liquid phase. We observe that the shape of the CO2 -dimer bands undergoes a distinctive line-mixing transformation, which reveals an unexpected stability of the dimers despite the collisions with the surrounding particles and negates the common belief that CO2 dimers are short-lived complexes. Our results furnish a deeper insight allowing a better modeling of CO2 -rich atmospheres and provide also a new spectroscopic tool for studying the robustness of molecular clusters.
Figure 1. Transmittances recorded in the ν1/2ν2 Fermi dyad region at 300 K. (a) Pure CO2 gas. Increasing gas pressure induces, simultaneously, a rise of the collision-induced absorption and a transformation of the dimeric bands. The gas pressures are converted into densities in amagat (Table II of Ref. 11). (b) 14.9-amagat CO2 mixed with various quantities of Ar. The absorption induced by collisions of CO2 monomers with Ar atoms has only a small influence on the band shape evolution which is dominated by the dimer signatures. [11] See supplementary material at http://dx.doi.org/ 10.1063/1.4906874 for experimental details, binary absorption coefficients, symmetric-top modeling, and pressure dependence of dimer band shape.
A Cavity Ring Down Spectrometer has been developed for high sensitivity absorption spectroscopy in the near infrared at pressure up to 10 bar. In order to strictly avoid perturbations of the optical alignment by pressure forces, the pre-aligned CRDS cavity is inserted inside the high pressure cell. The stability of the spectra baseline has been validated by filling the CRDS cell with Ar and N2 up to 10 bar.
We present here the first application of this CW-CRDS spectrometer to the study of the high pressure spectrum of CO2 at room temperature. The spectra were recorded between 5850 and 5960 cm−1 for a series of pressure values up to 6400 Torr. The studied spectral interval corresponds to the high energy range of the 1.75 µm transparency window of CO2 of particular interest for Venus.
The CO2 absorption coefficient at a given pressure value was obtained as the increase of the CRDS loss rate from its value at zero pressure taking into account the small contribution due to Rayleigh scattering. The CO2 absorption spectrum includes the contribution of the self broadened local rovibrational lines together with a broad and weak continuum. The CO2 continuum was obtained as the difference of the CO2 absorption coefficient and of a local lines simulation using the CO2 HITRAN line list and a truncated Voigt profile. A quadratic pressure dependence of the absorption continuum was observed, with an average binary absorption coefficient increasing smoothly from 3.7 to 11.5×10−9 cm−1 amagat−2 between 5850 and 5960 cm−1. The derived continuum shows a spectral feature located in the region of a band of 16O12C18O (present in natural abundance) which dominates the spectrum in the region. This spectral feature was found to arise from collisional interferences between local lines and quantitatively accounted for using a theoretical approach based on the impact and Energy Corrected Sudden (ECS) approximations. The impact of the chosen far wings CO2 line shape model on the retrieved continuum absorption is discussed.
Figure 5. Comparison between the measured CO2 absorption coefficients at 800, 1600, 3200 and 6400 Torr near 5930 cm-1 and the corresponding simulations of the absorption of the local rovibrational lines. The local line simulations used the HITRAN2012 line list and a Lorentzian line shape truncated at ±25 cm-1 (see Text).
Figure 3. Retrieval of the CO2 continuum absorption at 6400 Torr (lower panel) by difference between the absorption coefficient measured by CRDS and the corresponding simulations of the absortions of the local rovibrational lines.
Figure 3. mpact of the line mixing in the 10022-00001 band of 16O12C18O on the derived continuum absorption of CO2 (P=6400 Torr). Upper panel: Comparison to the CO2 CRDS spectrum at P = 6400 Torr of simulations of the absorption of the local rovibrational lines with and without taking into account line mixing effects, and corresponding difference. Lower panel: Resulting continuum absorption of CO2 obtained with and without taking into account line mixing effects.
The CO2 absorption was measured in the 7000 and 8000 cm–1 regions. The absorption coefficients were calculated using the asymptotic line wing shape theory. Line shape parameters were found from fitting to experimental data. The calculation results agree well with the measurement data. According to the line wing theory, the absorption in the band wings is caused by the wings of strong lines of an adjacent band. Within these assumptions, experimental and calculated data on the CO2 absorption coefficient in the band wings in the 7000 and 8000 cm–1 regions can provide information on the line shape at frequency detuning from several tens to several hundreds of half-widths. The results support the hypothesis that line shape parameters in the line wings related to transitions with the same initial state are close to each other. Deviations from a Lorentzian profile are found for some CO2 bands and turn out different for the wings of different bands
Figure 7. Absorption coefficient beyond the 4.3 μm band edge: (a) T = 193 K, experimental data [34] are marked by circles; (b) θ = 920 K, experimental data [35] are marked by circles; our calculation results are shown by the curves (see also [16–18]). [16] L. I. Nesmelova, O. B. Rodimova, and S. D. Tvorogov, Coefficient of light absorption in SO2 4.3 μm band. Izv. vuzov, Fiz., no. 10, 106–107 (1980). [17] L. I. Nesmelova, O. B. Rodimova, and S. D. Tvorogov, Spectral behavior of the absorption coefficients in the 4.3 μm CO2 band within a wide range of temperature and pressure, Atmos. Ocean. Opt. 5 (9), 609–614 (1992). [18] O. B. Rodimova, Spectral line profile of self-broadened CO2 from the center to the far wing, Atmos. Ocean. Opt. 15 (9), 694–703 (2002). [34] R. Le Doucen, C. Cousin, C. Boulet, and A. Henry, Temperature dependence of the absorption in the region beyond the 4.3 μmband of CO2. I: Pure CO2 case, Appl. Opt. 24 (6), 897–906 (1985). [35] J.M. Hartmann and C. Boulet, Line mixing and finite duration of collision effects in pure CO2 infrared spectra: Fitting and scaling analysis, J. Chem. Phys. 94 (10), 6406–6419 (1991).
Figure 3. Deviations from a Lorentzian profile for the 1.2, 1.2195, 1.4, 2.7, 4.3 μm CO2 bands calculated in (a) work [8] and (b) this work. The curve for 1.2195 μm almost coincides with the curve for 1.2 μm. [8] M. Y. Perrin and J. M. Hartmann, “Temperature_dependent measurements and modeling of absorption by CO2–N2 mixtures in the far line_wings of the 4.3 μm CO2 band,” J. Quant. Spectrosc. Radiat. Transfer 42 (4), 311–317 (1989)
The frequency and temperature dependence of the water vapor–nitrogen continuum in the 8–12 and 3–5 μm spectral regions obtained experimentally by CAVIAR and NIST is described with the use of the line contour constructed on the basis of asymptotic line shape theory. The parameters of the theory found from fitting the calculated values of the absorption coefficient to the pertinent experimental data enter into the expression for the classical potential describing the center-of-mass motion of interacting molecules and into the expression for the quantum potential of two interacting molecules. The frequency behavior of the line wing contours appears to depend on the band the lines of which make a major contribution to the absorption in a given spectral interval. The absorption coefficients in the wings of the band in question calculated with the line contours obtained for other bands are outside of experimental errors. The distinction in the line wing behavior may be explained by the difference in the quantum energies of molecules interacting in different vibrational states.
Figure 1. Н2О+N2 absorption coefficient in the 8–20 μm region. The full circles are the measured data from [7], the open circles are the calculated data from [19], and the curve is the calculation from [16]; Θ=296° (a)and 430°K (b).
[7] Burch DE, Alt RL. Continuum absorption by H2O in the 700–1200 cm-1 and 2400–2800 cm-1 windows. Scientific report no. 1. AFGL-TR-84-0128; 1984. [16] Ma Q, Tipping RH. The density matrix of H2O+N2 in the coordinate representation: a Monte Carlo calculation of the far-wing line shape. J Chem Phys 2000;112:574–84. [19] Rodimova OB. Theoretical researches of the absorption and photo-chemistry processes and their application in radiation codes of theclimate models [Thesis]. Dokt Phys-Math Sci Tomsk; 2003.
Слабоселективное (континуальное) поглощение электромагнитного излучения водяным паром является важным фактором, влияющим на радиационный баланс атмосферы Земли, а также – основным компонентом поглощения в ИК-окнах прозрачности атмосферы. Возможная физическая природа этого феномена дискутируется уже более 50 лет. В статье даются ретроспективный анализ и описание текущего состояния дел в решении проблемы, касающейся континуума водяного пара. Приводятся краткое описание существующих сегодня моделей континуального поглощения, их достоинства и недостатки, а также наиболее интересные экспериментальные и теоретические результаты последних лет, свидетельствующие о природе континуума.
Рисунок 2. Сглаженные спектры поглощения димера воды в миллиметровом (А) и в среднем ИК (Б) диапазонах согласно расчетам Викторовой и Жевакина (V&Zh) [12], Вигасина и Членовой (V&Ch) [14], Yukhnevich и Tarakanova (Yu&T) [15], Scribano и Leforestier (S&L) [16] и Lee и др. [17] в сравнении с экспериментальными данными Burch [22], Podobedov et al. [97] и полуэмпирической моделью континуума MT_CKD [25]. Рисунок заимствован из работы [4].
[4] Shine K.P., Ptashnik I.V., Rädel G. The water vapour continuum: Brief history and recent developments // Surv. Geophys. 2012. V. 33. Ð. 535–555. DOI: 10.1007/s10712-011-9170-y. [12] Викторова А.А., Жевакин С.А. Поглощение микрорадиоволн в воздухе димерами водяного пара // Докл. АН СССР. 1966. Т.171. №5. С.1061-1064. [14] Вигасин А.А., Членова Г.В. Спектр димеров воды в области длин волн > 8 мкм и ослабление излучения в атмосфере // Изв. АН СССР. Физ. атмосф. и океана. 1984. Т.20. №7. С.657-661. [15] Yukhnevich G.V., Tarakanova E.G. Some properties of the potential energy surface and vibrational spectrum of a strong hydrogen bond complex // J. Mol. Struct. 1988. V. 117. P. 495–512. [16] Scribano Y., Leforestier C. Contribution of water dimers absorption to the millimeter and far infrared atmospheric water continuum // J. Chem. Phys. 2007. V. 126. P. 234301-1–234301-12. [17] Lee M.-S., Baletto F., Kanhere D.G., Scandolo S. Far-infrared absorption of water clusters by firstprinciples molecular dynamics // J. Chem. Phys. 2008. V. 128. P. 214506-1–214506-5. [22] Burch D.E. Continuum absorption by H2O // Report AFGL-TR-81-0300. Air Force Geophysics Laboratory. Hanscom AFB, MA. 1981. 46 ð. [25] Mlawer E.J., Payne V.H., Moncet J.-L., Delamere J.S., Alvarado M.J., Tobin D.D. Development and recent evaluation of the MT_CKD model of continuum absorption // Phil. Trans. Roy. Soc. A. 2012. V. 370. Ð. 2520– 2556. DOI: 10.1098/rsta.2011.0295. [97] Podobedov V.B., Plusquellic D.F., Siegrist K.E., Fraser G.T., Ma Q., Tipping R.H. New measurements of the water vapor continuum in the region from 0.3 to 2.7 THz // J. Quant. Spectrosc. Radiat. Transfer. 2008. V. 109. P. 458–467.
Figure 9. Экспериментальный континуум чистого водяного пара из работы Paynter и др. [77] в полосах 1600 (а) и 3600 см-1(б), полученный при 295K, в сравнении с моделями спектра связанных (“Bound”) и квазисвязанных (“Quasibound”) димеров. Спектр связанных димеров смоделирован на основе данных об интенсивностях и центрах полос из работ [60,79], константы димеризации Kbeq=0,03 атм-1 и лоренцевского контура шириной 60 cm-1 для каждой полосы димера. Спектр квазисвязанных димеров смоделирован с использованием параметров линий мономера воды из HITRAN-2012 [80] с удвоенным интенсивностями, лоренцевской шириной 20 см-1 для каждой линии и константой димеризации Kmeq =0,06 атм-1. Суммарный спектр димеров показан жирной линией. Рисунок заимствован из работы [65].
[60] Kjaergaard H., Garden A., Chaban G., Gerber R., Matthews D., Stanton J. Calculation of vibrational transition frequencies and intensities in water dimer: Comparison of different vibrational approaches // J. Phys. Chem. A. 2008. V. 112. P. 4324–4335. [65] Ptashnik I.V., Shine K.P., Vigasin A.A. Water vapour self-continuum and water dimers. 1. Review and analysis of recent work // J. Quant. Spectrosc. Radiat. Transfer. 2011. V. 112. P. 1286–1303. [77] Paynter D.J., Ptashnik I.V., Shine K.P., Smith K.M., McPheat R., Williams R.G. Laboratory measurements of the water vapor continuum in the 1 200–8 000 cm–1 region between 293 and 351 K // J. Geophys. Res. 2009. V. 114. P. D21301-1–D21301-23. [79] Kuyanov-Prozument K., Choi M.Y., Vilesov A.F. Spectrum and infrared intensities of OH-stretching bands of water dimers // J. Chem. Phys. 2010. V. 132. P. 014304-1–014304-7. [80] Rothman L.S., Gordon I.E., Babikov I.E., Barbe A., Benner C.D., Bernath P.F., Birk M., Bizzocchi L., Boudon V., Brown L.R., Campargue A., Chance K., Cohen E.A., Coudert L.H., Devi V.M., Drouin B.J., Fayt A., Flaud J.-M., Gamache R.R., Harrison J.J., Hartmann J.-M., Hill C., Hodges J.T., Jacquemart D., Jolly A., Lamouroux J., Le Roy R.J., Li G., Long D.A., Lyulin O.M., Mackie C.J., Massie S.T., Mikhailenko S., Müller S.P., Naumenko O.V., Nikitin A.V., Orphal J., Perevalov V., Perrin A., Polovtseva E.R., Richard C., Smith M.A.H., Starikova E., Sung K., Tashkun S., Tennyson J., Toon G.C., Tyuterev Vl.G., Wagner G. The HITRAN 2012 molecular spectroscopic database // J. Quant. Spectrosc. Radiat. Transfer. 2013. V. 130. P. 4–50.
In this paper we investigate the water vapour continuum absorption retrieved from experimental high-resolution Fourier-Transform-Spectroscopy spectra, obtained earlier from high-temperature and pressure pure water vapour measuremets. The retrieved spectra of the water vapour continuum, related to the least-studied today near-IR absorption bands at 8800 and 10600 cm-1, are described on the basis of the water dimer hypothesis of the continuum. It is shown that water dimers should cause, at least, half of the water vapour continuum absorption in the studied spectral region. The temperature dependence of the effective dimerization constant is derived.
Figure 2. The results of fitting of the water dimers model spectrum to the retrieved continuum spectrum in the absorption bands of 8800 and 10600 cm-1 at a pressure of 1000 mbar, temperature of 400 K and an optical length of 17.7 m.
Spectral dependence of the self-continuum cross section at room temperature in the 1.25 μm window. Comparison of the MT_CKD2.8 model to the values obtained by CRDS-DFB and CRDS-ECDL. The blue curve corresponds to the recommended values provided as supporting information.
Spectroscopic catalogues, such as GEISA and HITRAN, do not yet include information on the water vapour continuum that pervades visible, infrared and microwave spectral regions. This is partly because, in some spectral regions, there are rather few laboratory measurements in conditions close to those in the Earth’s atmosphere; hence understanding of the characteristics of the continuum absorption is still emerging. This is particularly so in the near-infrared and visible, where there has been renewed interest and activity in recent years. In this paper we present a critical review focusing on recent laboratory measurements in two near-infrared window regions (centred on 4700 and 6300 cm−1) and include reference to the window centred on 2600 cm−1 where more measurements have been reported. The rather few available measurements, have used Fourier transform spectroscopy (FTS), cavity ring down spectroscopy, optical-feedback – cavity enhanced laser spectroscopy and, in very narrow regions, calorimetric interferometry. These systems have different advantages and disadvantages. Fourier Transform Spectroscopy can measure the continuum across both these and neighbouring windows; by contrast, the cavity laser techniques are limited to fewer wavenumbers, but have a much higher inherent sensitivity. The available results present a diverse view of the characteristics of continuum absorption, with differences in continuum strength exceeding a factor of 10 in the cores of these windows. In individual windows, the temperature dependence of the water vapour self-continuum differs significantly in the few sets of measurements that allow an analysis. The available data also indicate that the temperature dependence differs significantly between different near-infrared windows. These pioneering measurements provide an impetus for further measurements. Improvements and/or extensions in existing techniques would aid progress to a full characterisation of the continuum – as an example, we report pilot measurements of the water vapour self-continuum using a supercontinuum laser source coupled to an FTS. Such improvements, as well as additional measurements and analyses in other laboratories, would enable the inclusion of the water vapour continuum in future spectroscopic databases, and therefore allow for a more reliable forward modelling of the radiative properties of the atmosphere. It would also allow a more confident assessment of different theoretical descriptions of the underlying cause or causes of continuum absorption.
Figure 5. Temperature dependence of the water vapour self-continuum cross-section at three wavenumbers in the 4700 cm-1 window (after [58]) from Grenoble CRDS, OF-CEAS, CAVIAR FTS and Tomsk FTS, and the MT_CKD2.5 continuum model. The dashed lines show the slope of the curve assuming a temperature dependence of the form exp (D/kT) with Do = 1104 cm-1 (see text for details). (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)
[58] I. Ventrillard, D. Romanini, D. Mondelain, A. Campargue, Accurate measurements and temperature dependence of the water vapor selfcontinuum absorption in the 2.1μm atmospheric window, J. Chem. Phys. 143 (2015) 134304, http://dx.doi.org/10.1063/1.4931811. [59] D. Mondelain, S. Vasilchenko, P. Cermak, S. Kassi, A. Campargue, The self- and foreign-absorption continua of water vapor by cavity ring-down spectroscopy near 2.35 μm, Phys. Chem. Chem. Phys. 17 (2015) 17762–17770, http://dx.doi.org/10.1039/c5cp01238d. [23] I.V. Ptashnik, R.A. McPheat, K.P. Shine, K.M. Smith, R.G. Williams, Water vapor self-continuum absorption in near-infrared windows derived from laboratory measurements, Journal of Geophysical Research-Atmospheres, 116 (2011), D16305. 10.1029/2011JD015603. [55] I.V. Ptashnik, T.M. Petrova, Y.N. Ponomarev, K.P. Shine, A.A. Solodov, A.M. Solodov, Near-infrared water vapour self-continuum at close to room temperature, Journal of Quantitative Spectroscopy & Radiative Transfer, 120 (2013) 23-35. 10.1016/j.jqsrt.2013.02.016.
Figure 7. Temperature dependence of the water self-continuum cross sections, CS, from Grenoble CRDS [60], CAVIAR FTS [23], from Tomsk FTS [55] and Bicknell et al. CI [62] near the low energy edge (5875 cm-1), near the centre (6121 cm-1) and near the high energy edge (6665 cm-1) of the 6300 cm-1 window. Bicknell et al. reported only the sum of the self and foreign continua with a 30% error bar.
[23] I.V. Ptashnik, R.A. McPheat, K.P. Shine, K.M. Smith, R.G. Williams, Water vapor self-continuum absorption in near-infrared windows derived from laboratory measurements, Journal of Geophysical Research-Atmospheres, 116 (2011), D16305. 10.1029/2011JD015603. [55] I.V. Ptashnik, T.M. Petrova, Y.N. Ponomarev, K.P. Shine, A.A. Solodov, A.M. Solodov, Near-infrared water vapour self-continuum at close to room temperature, Journal of Quantitative Spectroscopy & Radiative Transfer, 120 (2013) 23-35. 10.1016/j.jqsrt.2013.02.016. [62] W.E. Bicknell, S.D. Cecca, M.K. Griffin, Search for low-absorption regimes in the 1.6 and 2.1 μm atmospheric windows, Journal of Directed Energy, 2 (2006) 151-161. [60] D. Mondelain, S. Manigand, S. Kassi, A. Campargue, Temperature dependence of the water vapor self-continuum by cavity ring-down spectroscopy in the 1.6 lm transparency window, J. Geophys. Res. – Atmos. 119 (2014) 5625–5639, http://dx.doi.org/10.1002/2013jd021319.
Figure 5. Spectral behavior of the water–carbon dioxide continuum in the 3–4 μm region. The data are a result of averaging over all four temperatures. The insert represents the temperature dependence of the binary absorption coefficients at three selected wavenumbers, designated in the plot by arrows. The bottom solid line shows the spectral behavior of the self-continuum [11] scaled roughly to the experimental data. [11] Baranov YuI. The continuum absorption in H2O-N2 mixtures in the 3 5 μm spectral region at temperatures from 326 to 363 K. J Quant Spectrosc Radiat Transf 2011;112:2281–6.
Figure 4. Spectral behavior of the water–carbon dioxide continuum in the 7 to 9 μm region. The data are a result of averaging over all four temperatures. The insert represents the temperature dependence of the binary absorption coefficients at four selected wavenumbers, designated in the plot by arrows. The thin solid line shows the spectral behavior of the MT_CKD [11] self-continuum fitted to the experimental poit at 1128 cm-1 and scaled to the experimental data/ The lower trace represents the pure CO2 spectrum in arbitrary units. [11] Baranov YuI. The continuum absorption in H2O-N2 mixtures in the 3 5 μm spectral region at temperatures from 326 to 363 K. J Quant Spectrosc Radiat Transf 2011;112:2281–6.
The very weak absorption continuum of CO2 is studied by Cavity Ring Down Spectroscopy in three 20 cm⁻¹ wide spectral intervals near the centre of the 1.74 µm window (5693 − 5795 cm⁻¹). For each spectral interval, a set of room temperature spectra is recorded at pressures between 0 and 10 bar thanks to a high pressure CRDS spectrometer. The absorption continuum is retrieved after subtraction of the contributions due to Rayleigh scattering and to local lines of CO2 and water (present as an impurity in the sample) from the measured extinction. Due to some deficiencies of the CO2 HITRAN2012 line list, a composite line list had to be built on the basis of the Ames calculated line list with line positions adjusted according to the Carbon Dioxide Spectroscopic Databank and self-broadening and pressure shift coefficients calculated with the Complex Robert Bonamy method. The local line contribution of the CO2 monomer is calculated using this list and a Voigt profile truncated at ±25 cm⁻¹ of the line centre. Line mixing effects were taken into account through the use of the impact and Energy Corrected Sudden approximations.
Figure 6. Impact of the line-mixing in the region of the R branch of the 00031–10002 band. Upper panel: CRDS spectrum, recorded for a density of 8.61 amagat (open circles) compared to the spectra simulations of the CO2 lines with (light grey) and without (red solid line) the line mixing effects. Lower panel: corresponding (Obs.–Sim.) residuals after subtraction of these contributions. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this.
The role of stable and metastable dimers as well as of free collisions in the collision-induced rototranslational absorption by the compressed CO2–Ar and CO2–Xe gas mixtures is elucidated using the classical three-dimensional trajectories method. The contribution from the stable dimers is obtained via Fourier transform of the dipole correlation function. The spectral bandshape due to the unbound trajectories (metastable dimers and free collisions) is calculated as an averaged Fourier spectrum of the collision-induced dipole moment. The mean lifetimes of metastable dimers have been estimated as 3.8 ps for CO2–Ar and 5.9 ps for CO2–Xe pairs. Trajectory computations are complemented by calculations of zero spectral moments using pair distribution functions. The stable and metastable dimer contribution to the zero spectral moment is shown to be comparable with that from free collisions.
Рисунок 7. Spectral functions for the CO2–Ar (a) and CO2–Xe (c) systems: experimental (dots) and calculated data (curves). Experimental data for the infrared region are taken from Ref. [33]. The Fig. 7(b) shows details of the spectral function in the microwave region with the experimental data obtained for the following wavenumbers: 0.8 [39], 2.3 [40], 4.6 and 15.1 cm-1 [41]. [33] Andreeva G.V, Kudriavtsev A.A., Tonkov M.V, Filippov N.N. Investigation of the integral characteristics off ar-IR absorption spectra of mixtures of CO2 with inert gases. Opt Spectrosc (USSR) 1990;68:623–5. [39] Maryott A.A., Kryder S.J. Collision-induced microwave absorption in compressed gases. III.CO2-foreign-Gas mixtures. J Chem Phys 1964;41:1580–2. [40] Dagg I.R., Reesor G.E. ,Urbaniak J.L.Collision-induced microwave-absorption in CO2 and CO2-Ar, CO2-CH4 mixtures in 2.3cm-1 region. Can JPhys1974;52:973–8. [41] Dagg I.R., Anderson A., Yan S., Smith W. The quadrupole moment of cyanogen: a comparative study of collision-induced absorption in gaseous C2N2, CO2, and mixtures with argon. Can J Phys 1986;64:1475–81.
Figure 5. Spectral function components for the CO2–Ar (a) and CO2–Xe systems (b) at 296 K. The curves show the contributions from collisions responsible for the formation of stable (1) and metastable (2) dimers and ordinary collisions (3).
Analysis of the continuum absorption in water vapor at room temperature within the purely rotational and fundamental ro-vibrational bands shows that a significant part (up to a half) of the observed absorption cannot be explained within the framework of the existing concepts of the continuum. Neither of the two most prominent mechanisms of continuum originating, namely, the far wings of monomer lines and the dimers, cannot reproduce the currently available experimental data adequately. We propose a new approach to developing a physically based model of the continuum. It is demonstrated that water dimers and wings of monomer lines may contribute equally to the continuum within the bands, and their contribution should be taken into account in the continuum model. We propose a physical mechanism giving missing justification for the super-Lorentzian behavior of the intermediate line wing. The qualitative validation of the proposed approach is given on the basis of a simple empirical model. The obtained results are directly indicative of the necessity to reconsider the existing line wing theory and can guide this consideration.
Figure 1. Left (a): χ-function calculated by Eq. (10) with the use of the parameters A=20 and Δνwing=11 cm-1 as a function of the detuning frequency. Right (b): Lorentz profile with a half-width of 5 *10-3 cm-1 (150 MHz) (bold solid line); the same profile with the wings cut off at the ±25 cm-1 detuning and the cut-off line “brought down” to the zero level (as in the CKD model) (thin solid line); the latter profile with the added wings, which are calculated by Eq. (8) using the χ-function shown on Fig.1a is presented by the dashed line.
Figure 2. Experimental spectra of the water vapor continuum bands at 1600 cm-1 from the paper17 approximated by the model including the contribution from the water dimers and the empirically calculated contribution from the line wings.The dots are the experimental data,the bold solid curve represents the model, the bold dashed curve, the thin dashed curve,and the thin solid curve are the contributions from the bound dimers,metastable dimers,and line wings, respectively. (a) The contribution from the line wings is calculated using Eqs. (8)–(10). A=32 for the 1600 cm-1 band , Δνwing=11 cm-1 for the both bands. (b) The contribution from the line wings is calculated using Eqs. (8)–(10). A=18.5 for the 3600 cm-1 band, Δνwing=11 cm-1 for the both bands.
[17.] Ptashnik I.V., Shine K.P., Vigasin A.A. Water vapour self-continuum and water dimers: 1. Analysis of recent work. J. Quant. Spectrosc. Radiat. Transf. 2011; 112: 1286–1303.>
Figure 3. Experimental spectra of the H2O continuum in far infrared [31] range approximated by the model allowing for the contribution from the water dimers and the empirically calculated contribution from the line wings.The dots are the experimental data,the bold solid curve represents the model,the bold dashed curve,the thin dashed curve,and the thin solid curve are the contributions from the bound dimers,metastable dimers,and line wings, respectively. (a)The contribution from the line wings calculated by using Eqs. (8)–(10). A=14 for the left-hand panel,and Δνwing=11 cm-1 in both cases. P=1 atm. T=296 K. (b) The contribution from the line wings calculated by using Eqs. (8)–(10). A=14 for the left-hand panel,and Δνwing=11 cm-1 in both cases. P=1 atm. T=296 K.
Figure 9. Spectral dependence of self-continuum cross-section at room temperature in the 2.1 μm window. MT_CKD3.0 model (purple solid line) is compared to (i) FTS results obtained in Tomsk [8, 9] or by the CAVIAR consortium [7], (ii) calorimetric–interferometry data [10] for which the blue open diamond corresponds to the self plus foreign cross-sections while blue full diamond is an approximated estimation of self-continuum cross-section value, (iii) our present and past CRDS [12, 14] and OF-CEAS [12, 13] measurements. The 30-50 % error bars on the Tomsk2015 FTS [9] values and the small error bars on our laser based values are not plotted for clarity. The insert shows the three CRDS measurement points of the present study. (Different symbols are used for the results obtained using upward and downward pressure ramps).
[7] Ptashnik I.V., McPheat R.A., Shine K.P., Smith K.M., Williams R.G. Water vapor self - continuum absorption in near - infrared windows derived from laboratory measurements. J Geophys Res 2011;116:D16305. doi: 10.1029/2011JD015603. [8] Ptashnik I.V., Petrova T.M., Ponomarev Y.N., Shine K.P., Solodov A.A., Solodov A.M. Near-infrared water vapour self-continuum at close to room temperature, J Quant Spectrosc Radiat Transfer 2013;120:23–35. doi: 10.1016/j.jqsrt.2013.02. 016. [9] Ptashnik I.V., Petrova T.M, Ponomarev Y.N., Solodov A.A., Solodov A.M. Water vapor continuum absorption in near-IR atmospheric windows. Atmos Oceanic Optics 2015;28:115–20. doi: 10.1134/S102485601502009. [10] Bicknell W.E., Cecca S.D., Griffin M.K . Search for low-absorption regimes in the 1.6 and 2.1 μm atmospheric windows. J Directed Energy 2006;2:151–61 . [12] Campargue A., Kassi S., Mondelain D., Vasilchenko S., Romanini D. Accurate laboratory determination of the near infrared water vapor self-continuum: A test of the MT_CKD model. J Geophys Res Atmos, 121,13,180–13,203, 2016, doi: 10.1002/ 2016JD025531. [13] Mondelain D., Vasilchenko S., Cermak P., Kassi S., Campargue A. The self- and foreign-absorption continua of water vapor by cavity ring-down spectroscopy near 2.35 μm. Phys Chem Chem Phys 2015;17(27) 17762-1777, 35 doi: 10.1039/ C5CP01238D . [14] Ventrillard I., Romanini D., Mondelain D., Campargue A.. Accurate measurements and temperature dependence of the water vapor self-continuum absorption in the 2.1 μm atmospheric window. J Chem Phys 2015;143. doi: 10.1063/1.4931811.
The amplitude, the temperature dependence, and the physical origin of the water vapour absorption continuum are a long-standing issue in molecular spectroscopy with direct impact in atmospheric and planetary sciences. In recent years, we have determined the self-continuum absorption of water vapour at different spectral points of the atmospheric windows at 4.0, 2.1, 1.6, and 1.25µm, by highly sensitive cavity-enhanced laser techniques. These accurate experimental constraints have been used to adjust the last version (3.2) of the semi-empirical MT_CKD model (Mlawer-Tobin_Clough-Kneizys-Davies), which is widely incorporated in atmospheric radiative-transfer codes. In the present work, the self-continuum cross-sections, CS, are newly determined at 3.3µm (3007cm−1) and 2.0µm (5000cm−1) by optical-feedback-cavity enhanced absorption spectroscopy (OFCEAS) and cavity ring-down spectroscopy (CRDS), respectively. These new data allow extending the spectral coverage of the 4.0 and 2.1µm windows, respectively, and testing the recently released 3.2 version of the MT_CKD continuum. By considering high temperature literature data together with our data, the temperature dependence of the self-continuum is also obtained.
Figure 6. Spectral dependence of self-continuum cross-section at room temperature in the 2.1 μm window (4200-5200 cm-1). The two last versions of the MT_CKD model (V3.0 and V3.2) are compared to the cavity-based measurements, to the FTS results obtained in Tomsk (Ptashnik et al., 2013, 2015) or by the CAVIAR consortium (Ptashnik et al., 2011). The 30–50% error bars on the Tomsk2015 FTS values and the small error bars on our laser based values are not plotted, for clarity. The zoom highlights the present CRDS values near 5000 cm-1.
Figure 7. Temperature dependence of the water vapour self-continuum cross-sections near 5006 cm-1 obtained by OFCEAS and CRDS (open and full red circles, respectively), by FTS (green squares: CAVIAR; full blue circles: Baranov and Lafferty, 2011; open blue squares: Tomsk 2013). The MT_CKD3.2 values which are normalised to the number density at 1 atm and 296°K were multiplied by 296/T . The D0 slope corresponds to an exp(D0=kT ) function, D0 ~ 1100 cm-1 being the dissociation energy of the water dimer molecule.
1. Ptashnik, I.V., McPheat, R.A., Shine, K.P., Smith, K. M., and Williams, R.G.: Water vapor self – continuum absorption in near – infrared windows derived from laboratory measurements, J. Geophys. Res., 116, D16305, 2011, https://doi.org/10.1029/2011JD015603. 2. Baranov, Y.I. and Lafferty, W. J.: The water-vapor continuum and selective absorption in the 3–5 µm spectral region at temperatures from 311° to 363°K, J. Quant. Spectrosc. Ra. Transf., 112, 1304–1313, 2011, https://doi.org/10.1016/j.jqsrt.2011.01.024. 3. Ptashnik, I. V., Petrova, T. M., Ponomarev, Y. N., Shine, K. P., Solodov, A. A., and Solodov, A. M.: Near-infrared water vapour self-continuum at close to room temperature, J. Quant. Spectrosc. Rad. Transf., 120, 23–35, 2013, https://doi.org/10.1016/j.jqsrt.2013.02.016.
Figure 5. Infrared absorption cross sections for the ν1 + ν6 band of methanol at a temperature of T = 298°K. The sample consisted of 202.2 μmol/mol of CH3OH-in-air at p equal to (a) 0.833 kPa, (b) 5.490 kPa, and (c) 26.859 kPa. Shown in (d) is a sample of 45.89 μmol/mol CH3OH-in-air at p = 101.575 kPa.
Figure 3. Comparison of the self-continuum optical depth retrieved within 1600 cm−1 (left panel) and 3600 cm−1 (right panel) absorption bands using HITRAN-2012 [26] with weak lines from UCL [27] and using HITRAN-2016 [31]. Measurement conditions: 11.5 mb pure water vapour at 288.5°K, path length 28.8 m.
[26] Rothman L., Gordon I., Babikov I., Barbe A., C. Benner D., et al. The HITRAN 2012 molecular spectroscopic database. JQSRT 2013;130:4–50. [27] Shillings A., Ball S., Barber M., Tennyson J., Jones R. An upper limit for water dimer absorption in the 750 nm spectral region and a revised water line list, Atmos Chem Phys 2011;10:23345–80. [31] Gordon I.E., Rothman L.S., Hill C., Kochanov R.V., Tan Y., Bernath P.F., Birk M., et al. The HITRAN 2016 molecular spectroscopic database. JQSRT 2017;203:3–69.
Figure 2. Example of pure water vapour continuum cross-section spectra within the 1600 and 3600 cm–1 bands, retrieved from high-resolution Fourier-transform spectra at below room temperatures [25] (solid circles) and at room and higher temperatures [24] (empty circles). Uncertainty in the retrieved continuum is shown for 288.5°К (15.35°C). The MT_CKD-3.2 self-continuum model [28] is also shown for comparison.
[24] Paynter D.J., Ptashnik I.V., Shine K.P., Smith K.M., McPheat R., Williams R.G. Laboratory measurements of the water vapor continuum in the 1200–80 00 cm–1 region between 293°K and 351K. J Geophys Res 2009;114:D21301 [25] Ptashnik I.V., Klimeshina T.E., Petrova T.M., Solodov A.A., Solodov A.M. Water vapor continuum absorption in the 2.7 and 6.25 μm bands at decreased temperatures. Atmos Oceanic Opt 2016;29(3):211–15. [28] Mlawer E., Payne V., Moncet J.L., Delamere M., Alvarado M., Tobin D. Development and evaluation of the MT_CKD model of continuum absorption. Philos Trans Royal Soc A 2012;370:2520–56.
Figure 5. Example of fitting of b- (green) + q-dimer (grey) model spectra to the experimental continuum, obtained in this work at different temperatures. The resulting total model (sum of b- and q-dimer) absorption is shown in red. For more accurate fitting, the wavenumbers of stable WD transition at 3597 and 3730 cm–1 [12] were shifted to 3616 and 3717 cm–1 respectively. The νi(PA) and νi(PD) on the top panels denote respectively proton-acceptor and proton-donor H2O unit in WD, while i = 1,2,3 mean correspondingly symmetric stretching, bending, and asymmetric stretching oscillations in the water unit. The values of Kq and Kb (in atm–1) that lead to a best fit are shown in the legends of each frame. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)
Figure 6. Equilibrium constants of the bound (left panel) and quasibound (right panel) water dimers derived in this work from fitting the model in Eq. (3) to experimental continua for two different absorption bands. Kb values retrieved from spectroscopic measurements in microwaves by Serov et al., theoretical calculations by Buryak and Vigasin, and modified Kb values from Scribano et al. calculations (see the text for details) are shown for comparison. The error bars represent the error in the experimental data on the derived continuum and error in fitting the model to the experimental data.
Figure 7. Total equilibrium constant (Kb+q), derived in this work. The data from Tretyakov et al. [44] , Ruscic [43] and Leforestier [42] are also shown for comparison. Thick grey line shows the ratio (right axis) of the total equilibrium constant derived from fitting in this work to that derived in [44] from SVC.
[42] Leforestier C. Water dimer equilibrium constant calculation: a quantum formulation including metastable states. J Chem Phys 2014;140:074106. [43] Ruscic B. Active thermochemical tables: water and water dimer. J Phys Chem A 2013;117(46):11940-53. [44] Tretyakov M.Yu., Serov E.A., Odintsova T.A. Equilibrium thermodynamic state of water vapour and the collisional interaction of molecules. Radiophys Quant Electron 2012;54(10):700-16